Next Article in Journal
Validation of a Liquid–Liquid Extraction Method to Study the Temporal Production of D-Series Resolvins by Head Kidney Cells from Atlantic Salmon (Salmon salar) Exposed to Docosahexaenoic Acid
Previous Article in Journal
Computation of Entropy Measures for Metal-Organic Frameworks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Boosted Tetracycline and Cr(VI) Simultaneous Cleanup over Z-Scheme WO3/CoO p-n Heterojunction with 0D/3D Structure under Visible Light

1
Hebei Province Collaborative Innovation Center for Sustainable Utilization of Water Resources and Optimization of Industrial Structure, Hebei Province Key Laboratory of Sustained Utilization & Development of Water Recourse, School of Water Resource and Environment, Hebei Geo University, Shijiazhuang 050031, China
2
School of Material Science and Engineering, Jiangsu University of Science and Technology, Zhenjiang 212003, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(12), 4727; https://doi.org/10.3390/molecules28124727
Submission received: 6 May 2023 / Revised: 2 June 2023 / Accepted: 8 June 2023 / Published: 13 June 2023
(This article belongs to the Section Photochemistry)

Abstract

:
In this study, a Z-Scheme WO3/CoO p-n heterojunction with a 0D/3D structure was designed and prepared via a simple solvothermal approach to remove the combined pollution of tetracycline and heavy metal Cr(VI) in water. The 0D WO3 nanoparticles adhered to the surface of the 3D octahedral CoO to facilitate the construction of Z-scheme p-n heterojunctions, which could avoid the deactivation of the monomeric material due to agglomeration, extend the optical response range, and separate the photogenerated electronhole pairs. The degradation efficiency of mixed pollutants after a 70 min reaction was significantly higher than that of monomeric TC and Cr(VI). Among them, a 70% WO3/CoO heterojunction had the best photocatalytic degradation effect on the mixture of TC and Cr(VI) pollutants, and the removing rate was 95.35% and 70.2%, respectively. Meanwhile, after five cycles, the removal rate of the mixed pollutants by the 70% WO3/CoO remained almost unchanged, indicating that the Z-scheme WO3/CoO p-n heterojunction has good stability. In addition, for an active component capture experiment, ESR and LC-MS were employed to reveal the possible Z-scheme pathway under the built-in electric field of the p-n heterojunction and photocatalytic removing mechanism of TC and Cr(VI). These results offer a promising idea for the treatment of the combined pollution of antibiotics and heavy metals by a Z-scheme WO3/CoO p-n heterojunction photocatalyst, and have broad application prospects: boosted tetracycline and Cr(VI) simultaneous cleanup over a Z-scheme WO3/CoO p-n heterojunction with a 0D/3D structure under visible light.

Graphical Abstract

1. Introduction

Since the 1940s, tetracycline (TC) has been recognized as an excellent antibiotic and has been widely used in human therapy, veterinary medicine, agriculture, and other fields [1,2,3]. However, the abuse of TC leads to its accumulation in the water environment [4]. Due to the fact that it is highly toxic, resistant, and difficult to biodegrade, TC accumulates in water for long periods, creating potential environmental stress [5,6,7]. In addition, heavy metals often coexist with antibiotic contaminants in water environments and cause irreversible damage to humans and other organisms [8,9,10]. The most typical is Cr (VI) contamination in water, which can be carcinogenic and cause lasting harm to the environment when ingested by humans [11,12]. Unfortunately, the combined toxicity of antibiotics and heavy metals is much higher than that of a single contaminant [13,14,15]. Antibiotic molecules contain a large number of carboxyl, hydroxyl, amino, heterocyclic, and other groups or electron donors, which can be complexed with heavy metal ions to change the environmental behavior and toxicological effects of pollutants in the complex pollution system, and may produce synergistic and antagonistic complex toxicological effects on micro-organisms, animals, and plants [16]. Therefore, seeking a new efficient, straightforward technology free of secondary pollution is essential for treating TC and Cr (VI) residues in water [17,18,19]. Semiconductor photocatalysis is an advanced redox technology that uses clean solar power as a power source to treat organic pollutants and heavy metal pollution in water [20,21,22]. Photocatalysis offers many advantages over conventional methods, including a simple preparation process, wide application range, good catalytic effect, and no secondary pollution [23,24], and it is considered by many researchers to be a superior and environmentally friendly purification technology [25,26,27,28].
Since Fujishu and Honda reported in 1972 that TiO2 has an excellent photocatalytic effect under ultraviolet light irradiation, traditional photocatalysis such as TiO2 and ZnO has been extensively studied [29]. However, TiO2 and ZnO have a wide band gap, extremely low visible light utilization, and fast electron–hole recombination, severely limiting their photocatalytic activity and effective utilization of visible light [30]. The p-type semiconductor cobalt oxide (CoO) is an excellent photocatalytic material with a smaller band gap of 2.2–2.6 eV, efficiently utilizing visible light [28,31]. However, the photocatalytic activity of pure CoO was limited by the high recombination rate of separate electron–hole pairs and inactivation due to agglomeration [32]. Fortunately, the construction of heterojunction nanocomposites can effectively solve the above problems and improve their photocatalytic efficiency [33,34]. In addition, three-dimensional octahedral CoO has the advantages of good stability, easy preparation, and good stability of its electrochemical reactions, and good prospects for the construction of heterojunctions [35,36,37]. For example, Zou et al. present a novel 3D/3D composite structure in which the 3D CoO acts as an active and stable ion diffusion channel. At the same time, the hollow NiCo LDH provides abundant redox sites. Chen et al. [38] reported the oxygen-vacancy-induced construction of the CoO/h-TiO2 Z-scheme heterostructures, which can be used for photocatalytic hydrogen production from water decomposition. Wang et al. [39] designed a novel one-dimensional/two-dimensional (1D/2D) core-shell cobalt monoxide/nickel-cobalt layered double hydroxide (CoO/Nico-LDH) heterojunction, which can avoid agglomeration, promote the transfer of photogenerated carriers, and expand the light absorption range, effectively enhancing the photocatalytic performance. As a conventional photocatalytic material, WO3 has a band gap of about 2.5–2.8 eV [40], can absorb visible light at nearly 500 nm, and has good photocatalytic performance in visible light [41,42,43]. Cao et al. [44] reported the electron transfer mechanism of noble-metal-free WO3@ZnIn2S4 S-scheme heterojunction photocatalysts. This provides a new idea for improving the photocatalytic efficiency of WO3 through the formation of heterojunctions [45]. Consequently, Z-scheme p-n heterojunctions can be constructed using WO3 as 0D nanoparticles attached to 3D octahedral CoO [46,47,48]. This can not only broaden the visible light response range of the monomeric material but also effectively improve the separation efficiency of the photogenerated carriers and the stability of the material [49,50,51].
In this article, Z-scheme WO3/CoO p-n heterojunctions were designed and synthesized via a simple solvothermal approach. The XRD, TEM, XPS, and UV-vis measurement were chosen to characterize the microstructure, surface chemistry, and photoelectrochemical properties of WO3/CoO p-n heterojunctions. In addition, the reusability and stability of the Z-scheme WO3/CoO p-n heterojunctions were tested by five cycles. Finally, the potential photocatalytic mechanisms in the photocatalytic degradation of TC were elaborated.

2. Results and Discussion

2.1. XRD Analysis

X-ray diffraction (XRD) was used to study the phase structure of the as-synthesized WO3, CoO, and various WO3/CoO heterojunctions. As indicated in Figure 1, the diffraction peaks of pure WO3 at 2θ = 23.12°, 23.59°, 24.38°, 33.26°, 33.57°, 34.15°, and 49.95° corresponded to the (002), (020), (200), (022), (−202), (202), and (140) crystal planes of the monoclinic phase (JCPDS card NO. 43-1035), respectively [52,53]. Moreover, the characteristic peaks at 36.48° and 42.37° in pure CoO and different WO3/CoO heterojunctions are normalized to the (111) and (200) planes of cubic phase CoO (JCPDS 71-1178) [54,55]. With the construction of the WO3/CoO heterojunction, the characteristic peak of WO3 increased with the increase of content, and no other obvious peaks were found, indicating that no impurities were observed during the recombination process. It was worth noting that the prominent diffraction peaks of WO3 and CoO had slightly shifted to the middle, which can be ascribed to the metal ion bonding in creating heterojunctions between WO3 and CoO.

2.2. Morphology

SEM was employed to observe the morphology of the WO3, CoO, and 70% WO3/CoO heterojunction, and shown in Figure S1. As illustrated in Figure S1a, CoO possessed an octahedral form with a smooth surface and an average size of about 200 nm. WO3 appeared as a nanosheet structure of agglomeration with a side length of about 350 nm. In Figure S1c, the morphology and particle size of the WO3/CoO heterojunction was very similar to that of pure CoO, indicating that the addition of WO3 had no significant change in the morphology and particle size of CoO. After the combination of WO3 and CoO, WO3 nanoparticles adhered to the surface of the octahedral CoO, indicating that WO3/CoO heterojunction was well-constructed, whereafter the specific morphology of the WO3, CoO, and the 70% WO3/CoO heterojunction was further analyzed by TEM. Figure 2a clearly showed that the octahedral CoO was stacked together and the size was in accord with the above SEM results. As shown in Figure 2b, WO3 presented the nanosheet structure with accumulation and in agreement with the morphology of SEM. As shown in Figure 2c,d, it can be seen that the CoO composite material still presented a representative octahedral shape, and WO3 changed from sheet to granular during preparation and deposited on the surface of the octahedral CoO. Figure 2e showed the HR-TEM images of the 70% WO3/CoO heterojunction material for a lattice study. It can be seen from the figure that the 0.25 nm lattice fringe spacing corresponded to the (111) plane of the CoO composite material based on JCPDS 71-1178, while the 0.21 nm lattice spacing corresponded to the (202) plane of the WO3 material according to the standard card (JCPDS 43-1035) [52,55].

2.3. XPS

The surface chemical states, chemical composition, and molecular structure of the as-prepared photocatalysts were examined by XPS spectroscopy. As shown in Figure S2, the XPS survey spectrum of 70%WO3/CoO consists of Co, W, and O, with all characteristic peaks of WO3 and CoO, and no impurities other than carbon were found in the spectrum element, consistent with the XRD results. The C1s peak observed at 284.4 eV can be attributed to the signal from carbon during the subsequent processing of the measurement of the scanning spectrum. Figure 3a displays the high-resolution Co 2p spectrum. As shown in the figure, the two main peaks were located at 780.2 and 796.3 eV, while the two satellite peaks were located near 786.2 eV and 802.9 eV, which correspond to Co 2p1/2 and Co 2p3/2, manifesting the occurrence of Co2+ [56]. As depicted in Figure 3b, two peaks of 35.4 eV and 37.4 eV are attributed to W4 f7/2 and W4 f5/2 in WO3, respectively, belonging to the W6+ in WO3. After the combination of CoO and WO3, the W 4f7/2 and W 4f5/2 of WO3/CoO drift to 35.2 eV and 37.0 eV, which were slightly lower than the W 4f7/2 and W 4f5/2 of pure WO3 [57]. On the contrary, the Co 2p peaks of WO3/CoO were slightly higher than those of pure CoO. The above results indicated that electrons were transferred from the CB of WO3 to the VB of CoO with the following Z-scheme pathway, meaning the Z-scheme WO3/CoO heterostructure had been successfully constructed [58,59]. Furthermore, the O 1s spectrum of WO3/CoO could be resolved as triple peaks at 530.0 eV, 531.0 eV, and 532.0 eV, as shown in Figure 3d. Among them, the formation of the characteristic peak at 530.0 eV was due to the typical lattice oxygen in CoO and WO3 [60,61], while the characteristic peaks were also observed at 531.0 eV and 532.0 eV, possibly due to the combination of O in H2O on the surface of WO3/CoO with active species (•OH and •O2) [62,63].

2.4. UV-Vis

UV-Vis diffuse reflectance spectroscopy was employed to investigate the optical absorption characteristics, band gaps, and energy levels of the CoO, WO3, and WO3/CoO heterojunctions. In Figure 4a, since pure WO3 had a narrow band gap width of 2.9 eV, it had an absorption edge at approximately 466 nm and strong light absorption properties in both the UV and visible spectra [42,45,48]. However, the light absorption intensity of CoO decreased significantly with increasing wavelength, which limited the photocatalytic performance, which was also consistent with the previous studies. The difference was that pure CoO could effectively use UV and visible light [64]. Compared with pure WO3, the absorption spectra of the WO3/CoO heterojunctions showed a red shift and a significant enhancement of the absorption band edge, indicating that the constructed WO3/CoO heterojunctions could effectively enhance the light absorption performance and thus improve the photocatalytic activity [31,41]. As shown in Figure 4b, the forbidden bandwidths of pure WO3 and CoO were 2.90 eV and 2.34 eV, respectively, as calculated by the Tauc plot equation (Formula 1). Meanwhile, as shown in Figure 5c,d, the flat-band potentials of pure WO3 and CoO were determined using electrochemical Mott–Schottky and displayed at −0.45 V and 0.87 V (vs. NHE), respectively. Due to the negative slope of CoO and positive slope of WO3, CoO and WO3 were p-type and n-type semiconductors, respectively. Therefore, we can deduce that the potentials of the VB of CoO and the CB of WO3 were 0.97 eV and −0.55 eV (vs. NHE), respectively. In summary, the WO3 VB potential was 2.35 eV, and the CoO CB potential was −1.37 eV (vs. NHE). In conclusion, it is proven that CoO and WO3 could construct Z-scheme WO3/CoO p-n heterojunctions in this situation.
A h ν = A ( h ν - Eg ) 1 / n
α, h, V, Eg, and A are the absorption coefficient, Planck’s constant, incident light frequency, band gap energy, and constant, respectively. The value of N determines the characteristics of transitions in semiconductors, with values equal to 1/2 and 2 representing indirect and direct transitions, respectively.

2.5. Photocatalytic Activites

Figure 5 illustrated the photocatalytic performance of pure CoO, pure WO3, and 10–90% WO3/CoO heterojunctions for the degradation of two monomeric pollutants of TC and Cr(VI) and a mixture of TC and Cr(VI) under the illumination of visible light. Under natural conditions, only 5% of the TC was degraded, as shown in Figure 5a, indicating that it was difficult to achieve the self-degradation of TC in water, while the increase of light time after the photocatalyst injection significantly decreased the concentration of TC. The pure CoO and WO3 degraded 38.7% and 34.2% of the TC after 70 min of the photocatalytic reaction. In contrast, 10–90% of the WO3/CoO degraded 70.7%, 76.1%, 82.7%, 93.8%, and 81.6% of the TC, respectively. This proved that constructing WO3/CoO heterojunctions could effectively improve photocatalytic activity. Among them, 70% WO3/CoO heterojunctions exhibited the best performance of the photocatalyst. As shown in Figure 5b, based on the Langmuir–Hinshelwood kinetic model, the kinetic curve had good linear characteristics, indicating that the photocatalytic oxidation of TC follows the quasi-first-order kinetic model. In Figure 5c, it can be seen that, during the degradation of TC, the 70% WO3/CoO reaction rate constant is 0.0386 min−1, which is 6.26 times and 6.82 times higher than pure CoO (0.00617 min−1) and WO3 (0.00566 min−1), further proving that 70% WO3/CoO has the best photocatalytic degradation efficiency for TC. As shown in Figure 5d, the removal of Cr(VI) by pure CoO, WO3, and 10–90% WO3/CoO after 70 min of reaction was 18.7%, 37.4%, 40.0%, 44.1%, 34.7%, 31.0%, and 21.8%, respectively. Figure 5e shows that the kinetic curve exhibits good linear characteristics, indicating that the photocatalytic removal of Cr (VI) follows a quasi-first-order kinetic model. For other mass ratios of WO3/CoO, the reaction rate constant (0.00819 min−1) of 70% WO3/CoO(Figure 5f) is still the highest, which is 2.95 times and 2.43 times higher than pure CoO (0.00278 min−1) and WO3 (0.00337 min−1), respectively. Considering the coexistence of TC and Cr(VI) in real wastewater, the photocatalytic simultaneous cleanup of TC and Cr(VI) in the mixed solution were measured. Figure 5g demonstrated that the degradation efficiency of pollutants mixed after a 70 min reaction was significantly greater than that of a single TC and Cr(VI). Figure 5h shows that removing TC and Cr (VI) mixed pollutants also follows a quasi-first-order kinetic model. The degradation rate of TC and Cr(VI) increased from 93.8% and 44.13% to 95.35% and 70.2%, respectively, after five cycles, demonstrating that the WO3/CoO heterojunction had excellent stability (Figure 5i).

2.6. Electrochemical Test

To further investigate and explore the reasons for the improved photocatalytic performance of WO3/CoO heterojunctions, the photocurrent- response, electrochemical impedance, photoluminescence spectra, and time-resolved photoluminescence decay curves were used to probe the carrier capture, migration, and electron–hole pair separation efficiency. In Figure 6a,b, the apparent current density of 70% WO3/CoO was greater than pure CoO and WO3, while the EIS arc radius of 70% WO3/CoO arc was the smallest. The above results showed that the construction of the WO3/CoO heterojunction can significantly improve the separation efficiency and charge the transfer capability of photogenerated carriers. Subsequently, the 70% WO3/CoO heterojunction showed the lowest PL intensity peak in Figure 6c, manifesting that, after successfully constructing the WO3/CoO p-n heterostructure, the recombination of electron–hole pairs was effectively suppressed. Moreover, PL lifetimes of photogenerated electron–hole pairs were determined by time-resolved phosphor spectroscopy. The 70% WO3/CoO heterojunction revealed a longer lifetime than those of pure CoO and WO3 (Figure 6d), which further indicated that the construction of WO3/CoO heterojunctions can effectively extend the lifetime of photogenerated carriers and thus restrain the recombination of photogenerated e and h+, which was conducive to improving the photocatalytic performance.

2.7. Radical Trapping and ESR

Active species capture and ESR experiments were performed for determining the significant active species involved in the reaction system and their contribution. EDTA-2Na, BQ, and IPA were selected as scavengers. As shown in Figure 7a, the TC degradation rate reduced from 93.8% to 24.5%, 29.2%, and 36.7%, respectively, indicating that e and •O2 contributed significantly in the reaction, while •OH was also involved in the photocatalytic reaction. To further verify the existence of free radicals, the ESR analysis was performed under visible light. As shown in Figure 7b, no ESR signal regular of DMPO-•O2 and DMPO-•OH was observed under dark conditions. However, a series of obvious characteristic peaks was received under light conditions, indicating that •O2 and •OH free radicals were involved and played critical roles in photocatalytic reactions.

2.8. LC-MS

Liquid chromatography–mass spectrometry (LC-MS) was used to validate intermediates to further determine the photocatalytic degradation pathway of TC. The mass spectrum resulting from the degradation of the original TC and corresponding intermediates after 70 min was shown in Figure S3. The protonated tetracycline molecule was represented by the single peak of the original TC sample at m/z = 445 [65,66,67,68,69]. The degradation of TC into many intermediates was observed after 70 min when the MS peak at m/z = 445 was significantly reduced, and several new peaks appeared at m/z = 416, 318, 279, 218, 173, 150, 118, and 111. Based on the combination of the previous studies and the above mass spectra, the structural information of the potential intermediates produced in the photocatalytic process was presented in Table S1. Therefore, Figure 8 shows the potential pathways of TC degradation. Firstly, TC lost two methyl groups to obtain P1 (m/z = 416). Following this, there are three pathways in further transformation. For Pathway 1, P2 (m/z = 279) was produced from P1 (m/z = 416) by ring opening and decarboxylate, and the formation of P3 (m/z = 173) was formed by ring opening and demethylation from P2 (m/z = 279), then P4 (m/z = 118) was achieved through product P3 by ring opening. For Pathway 2, P5 (m/z= 318) was generated from product P1 (m/z = 416) by the cleavage of the double bond and the removal of the carboxyl group and amino group. Following this, P6 (m/z = 218) was converted from P5 (m/z= 318) via ring opening, deamination, and double-bond oxygen. Subsequently, P7 (m/z = 149) was obtained by ring opening. For the third pathway, the product P8 (m/z = 318) was converted from P1 by dehydroxylation, deaminationring, cleavage of bond, and ring opening. Then, the product P9 (m/z = 218) was obtained by the oxidation of the double bond to the single bond and dehydroxylation. Subsequently, product P10 (m/z = 111) was converted from product P9 by dehydroxylation, double-bond cleavage, and ring opening. In summary, P4, P7, and P10 were partially mineralized into CO2, H2O, NH4+.

2.9. Photocatalytic Mechanism

On the basis of the above analysis and results, the degradation mechanism of the TC and Cr(VI) of the Z-scheme WO3/CoO p-n heterojunction under visible light was proposed with the following, as shown in Figure 9. In terms of the Mott–Schottky diagram, CB and VB are shown in Figure 9a. It is clear that the band gap positions of WO3 and CoO can form type II heterojunctions [70,71]. Electrons transferred to the CB of WO3 and CoO, and h+ was generated at the VB of WO3 and CoO when WO3/CoO was irradiated with visible light. Because a p-n heterojunction was formed, photogenerated electrons from the CB of CoO were transferred to the CB of WO3 under the influence of the potential difference, while h+ from the VB of WO3 could be transferred to the VB of CoO. However, the CB potential of WO3 (−0.55 eV vs. NHE) is smaller than that of the superoxide radical O2/•O2(−0.33 eV vs. NHE), and the VB potential of CoO (0.97 eV vs. NHE) is smaller than that of the hydroxyl radical (E(•OH/OH) = +1.99 eV vs. NHE). [6,72,73,74]. This indicated that •O2 could be produced during photocatalysis without •OH, which apparently contradicted the ESR results. Therefore, based on the above results and previous literature, a p-n heterojunction that follows the Z-scheme path mechanism was proposed. As shown in Figure 9b, when the p-type CoO made contact with the n-type WO3, the electrons on WO3 diffused to the CoO surface and suppressed the agglomeration of CoO. At the same time, the Fermi energy of the WO3/CoO tended to be balanced due to the electric field created among WO3 and CoO. Under visible light irradiation, photogenerated electrons migrated from the CB of WO3 to the VB of CoO according to the Z-scheme pathway, preserving the significant redox currents of the CB of CoO and the VB of WO3 [75,76,77]. In this case, the main active species produced by the Z-scheme mechanism are •O2 and •OH, in agreement with the ESR results. TC decomposes into small molecule substances such as CO2, H2O, and NH4+ under the synergistic action of •O2, •OH, and h+ [5,51,78,79]. Under the act of e and •O2 at the CB of WO3, Cr(VI) is reduced to Cr(III). In addition, e plays a significant role as an electron donor in the photocatalytic reduction of Cr(VI), and the photocatalytic degradation of TC can effectively consume excessive h+, thus inhibiting the recombination of the photogenerated e and h+, and ultimately further improving the photocatalytic efficiency of the synergistic removal of Cr (VI) and TC [80,81]. In summary, the Z-scheme WO3/CoO p-n heterojunctions could effectively facilitate the separation of photogenerated carriers and improve photocatalytic activity.

3. Materials and Methods

3.1. Chemicals and Materials

Anhydrous ethanol is purchased from Tianjin Yongda Chemical Reagent Co., Ltd. (Tianjin, China). Sodium tungstate, N-octanol, and cobalt acetate are supplied by Tianjin Damao Chemical Reagent Factory (Tianjin, China), and tetracycline is purchased from Shanghai Baoman Biotechnology Co (Shanghai, China). Nitric acid was purchased from China National Pharmaceutical Group Chemical Reagent Co., Ltd. (Shanghai, China). All experimental water is pure water, and the chemical substances used in the experiment, except for nitric acid, are analytically pure and can be used without further purification.

3.2. Characterization

The morphology was evaluated by scanning electron microscopy (SEM) with Phenom ProX, and the microstructure and elements of the composites were analyzed. The transmission electron microscope (TEM) photographs were taken using a transmission electron microscope (FEI, Hillsboro, USA) to observe the structural characteristics of the composites. X-ray diffraction (XRD) of the specimens was performed with Pert-ProMPD/max-γAX-ray (Cu Ka radiation with λ = 1.5406A, 2θ: 10–80°). X-ray photoelectron spectroscopy (XPS) was performed on an Axis ultra-DLD photoelectron spectrometer (Manchester, Britain) to analyze the nature, content, valence, and state of the composite. A Lambda 750 (PerkinElmer, Boston, MA, USA) spectrometer was used to obtain UV-Vis absorption spectra to reflect the optical absorption properties. CoO and WO3 were used as a reference to discuss the material’s electronic structure and the semiconductor’s band gap.

3.3. Synthesis of WO3, CoO, and WO3/CoO Heterojunctions

First, 5 mL of 65% concentrated nitric acid (HNO3) was slowly added into 25 mL of deionized water, thus obtaining diluted nitric acid. Then, 0.519 g of sodium tungstate (Na2WO4·2H2O) was dissolved in 10 mL of deionized water, and stirred until Na2WO4·2H2O dissolved. Then, as-prepared diluted nitric acid was slowly added into above Na2WO4·2H2O aqueous solution, and ultrasonic-stirred for 30 min. This process changed the sediment from white to faint yellow. The resulting mixture was added to a 50 mL Teflon-lined stainless-steel reactor and heated at a constant temperature of 180 °C for a period of 12 h. At the end of the reaction, the samples were naturally cooled to room temperature, washed alternately with deionized water and ethanol, and placed in a drying oven at 80 °C for 6 h to obtain WO3.
The WO3/CoO heterojunction were prepared by facile and robust solvothermal synthesis as follows: Take the 90 wt.% WO3/CoO heterojunction, for example: 1.84 g cobalt acetate (Co(CH3COO)2·4H2O) was dissolved into the mixture solution of 64 mL octyl alcohol and 16 mL ethanol and stirred for 2 h at room temperature. Then, we added 0.1466 g WO3 into the above mixture solution and stirred magnetically for 10 min. Then, the suspension was poured into 100 mL Teflon-lined stainless-steel reactor and heated in oven at 220 °C for 4 h. At the end of the reaction, the samples were naturally cooled to room temperature, obtained by centrifugation, and washed three times alternately with deionized water and ethanol, and finally taken to vacuum-drying oven at 70 °C for 6 h to obtain 90 wt% WO3/CoO heterojunctions. Finally, WO3/CoO heterojunctions with various mass ratios were produced by adding the corresponding mass ratio of WO3 to the above mixed solution under the same conditions and labeled as 10% WO3/CoO, 30% WO3/CoO, 50% WO3/CoO, 70% WO3/CoO, and 90% WO3/CoO. The fabrication scheme of the WO3/CoO p-n heterojunction photocatalyst is presented in Figure 10.

3.4. Measurement of Photocatalytic Activity

We separately added 30 mg of 10–90% WO3/CoO to the TC solution (50 mL, 40 mg/L) and stirred for 30 min under dark conditions to ensure that the TC reaches adsorption equilibrium on the photocatalyst surface. A group of pure TC solution without any photocatalyst was set as blank control, and its initial concentration was denoted as C0. A 300 W xenon lamp with a λ > 420 nm cut filter was used as a visible light source to irradiation solution, with continuous magnetic stirring at atmospheric pressure, and 3 mL of the suspension was taken at 15 min interval and centrifuged to remove the photo catalyst. The experimental temperature was kept at 25 °C to eliminate the effect of temperature on the experiment. The supernatant was measured by setting the maximum wavelength of the UV-Vis spectrophotometer to 357 nm. The degradation rate of the TC solution can be calculated using the following equation:
D e g r a d a t i o n % = 1 - C C 0 × 100 %
C0 is the absorbance of the starting solvent, and C is the absorbance of the solvent after the reaction. In order to test the stability of photocatalytic per form, the sample with the best photocatalytic activity were selected for five cycles of experiment; that is, the solution was recovered and centrifuged after the experiment, and cleaned with ethanol and water for several times and dried again to obtain the available photocatalyst, and then the degradation rate of TC was tested using the same method.
Similar to the photocatalytic degradation of the TC experiment, Cr (VI) solution (30 mg/L) was used as the target pollutant. During the irradiation process, 3 mL of the supernatant was taken out every 15 min, and the concentration of Cr (VI) was analyzed at 540 nm using a UV-visible spectrophotometer.

3.5. Reactive Radical Trapping Experiments

Similar to the above photocatalysis experiment, keep the original experimental process unchanged. Before the reaction, 1 mmol ethylenediaminetetraacetic acid disodium salt (EDTA-2Na), 1,4-Benzoquinone (BQ), and isopropanol (IPA) captors are added to the TC solution to explore the contribution of superoxide anion radical (O2), hydroxyl radical (OH), and hole (h+) in the process of photocatalysis experiment.

4. Conclusions

In conclusion, Z-scheme WO3/CoO p-n heterojunction photocatalysts were successfully prepared by a simple solvothermal method. XRD, SEM, TEM and UV-Vis were used to study the morphology, crystal structure, composition, optical properties, and carrier transfer mechanism. The results show that TC and Cr(VI) can be effectively dislodged by the 70% WO3/CoO heterojunction photocatalyst. At the same time, after five cycles, the removal rate of the TC and Cr(VI) of the WO3/CoO p-n heterojunction was only slightly reduced, indicating that the heterojunction had excellent stability. Moreover, the 70% WO3/CoO heterojunction photocatalyst showed higher photocatalytic activity than pure CoO and WO3, and the reaction rate constants of TC degradation are 6.26 times and 6.82 times that of pure CoO and WO3, respectively, while the reaction rate constants of Cr(VI) degradation are 2.95 times and 2.43 times that of pure CoO and WO3, respectively. Finally, the potential route and photocatalytic process of the degradation of TC and Cr(VI) were determined by an active species trapping experiment, ESR, and LC-ms spectrometry. In summary, new perspectives are provided by these findings in the construction of high-performance photocatalysts for p-n heterojunctions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28124727/s1, Figure S1 SEM images of (a) CoO, (b) WO3, and (c) 70% CoO-WO3. Survey spectra of 70% WO3/CoO heterojunction. Figure S3 Mass spectra of doxycycline hydrochloride of 70% WO3/CoO. Figure S4 Nitrogen adsorption desorption isotherm and pore size distribution curve (a–e) 10–90% WO3/CoO, (f) CoO, (g) WO3; Dark adsorption experiment (h) TC, (i) Cr (VI). Figure S5 (a,b) TEM and HRTEM images of 70% WO3/CoO; (c–f) EDX mapping of 70% WO3/CoO. Figure S6 (a) TEM images, (b) XRD of 70% WO3/CoO after the photocatalytic process. Figure S7 10–90% WO3/CoO composite material bandgap. Table S1 Possible intermediate products [82,83,84,85].

Author Contributions

C.L.: conceptualization, funding acquisition, and writing—review & editing. D.C.: writing—original draft, investigation, and data curation. H.Z.: conceptualization, investigation, and formal analysis. L.G.: investigation and formal analysis. W.S.: funding acquisition and writing—review & editing. F.G.: formal analysis and funding acquisition. Y.Z.: methodology and review & editing. J.L.: validation, project administration, supervision, writing—review & editing, and funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the National Natural Science Foundation of China (21906039, and 22006057), Hebei Province 333 Talents Project (A202101020), Science and Technology Project of Hebei Education Department (BJ2021010), Graduate Student Innovation Ability Training Funding Project of Hebei Province (CXZZSS2023129), Open Fund for Hebei Province Key Laboratory of Sustained Utilization & Development of Water Recourse (HSZYL2022002), Hebei Geo University Student Science and Technology Fund (KAG202306), National Pre-research Funds of Hebei GEO University in 2023 (Grant2022015), and 2022 Hebei GEO University Undergraduate Innovation and Entrepreneurship Training Program (X202210077036).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data of the study can be provided by the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Hu, Y.; Chen, D.Z.; Zhang, R.; Ding, Y.; Ren, Z.; Fu, M.S.; Cao, X.K.; Zeng, G.S. Singlet oxygen-dominated activation of peroxymonosulfate by passion fruit shell derived biochar for catalytic degradation of tetracycline through a non-radical oxidation pathway. J. Hazard. Mater. 2021, 419, 126495. [Google Scholar] [CrossRef] [PubMed]
  2. Li, S.J.; Wang, C.C.; Liu, Y.P.; Cai, M.J.; Wang, Y.N.; Zhang, H.Q.; Guo, Y.; Zhao, W.; Wang, Z.H.; Chen, X.B. Photocatalytic degradation of tetracycline antibiotic by a novel Bi2Sn2O7/Bi2MoO6 S-scheme heterojunction: Performance, mechanism insight and toxicity assessment. Chem. Eng. J. 2022, 429, 132519. [Google Scholar] [CrossRef]
  3. Wang, C.C.; Cai, M.J.; Liu, Y.P.; Yang, F.; Zhang, H.Q.; Liu, J.S.; Li, S.J. Facile construction of novel organic-inorganic tetra (4-carboxyphenyl) porphyrin/Bi2MoO6 heterojunction for tetracycline degradation: Performance, degradation pathways, intermediate toxicity analysis and mechanism insight. J. Colloid Interface Sci. 2022, 605, 727–740. [Google Scholar] [CrossRef] [PubMed]
  4. Luo, J.; Lin, P.P.; Zheng, P.L.; Zhou, X.F.; Ning, X.M.; Zhan, L.; Wu, Z.J.; Liu, X.N.; Zhou, X.S. In suit constructing S-scheme FeOOH/MgIn2S4 heterojunction with boosted interfacial charge separation and redox activity for efficiently eliminating antibiotic pollutant. Chemosphere 2022, 298, 134297. [Google Scholar] [CrossRef]
  5. Li, S.J.; Chen, J.L.; Hu, S.W.; Wang, H.L.; Jiang, W.; Chen, X.B. Facile construction of novel Bi2WO6/Ta3N5 Z-scheme heterojunction nanofibers for efficient degradation of harmful pharmaceutical pollutants. Chem. Eng. J. 2020, 402, 126165. [Google Scholar] [CrossRef]
  6. Li, S.J.; Wang, C.C.; Liu, Y.P.; Xue, B.; Jiang, W.; Liu, Y.; Mo, L.Y.; Chen, X.B. Photocatalytic degradation of antibiotics using a novel Ag/Ag2S/Bi2MoO6 plasmonic p-n heterojunction photocatalyst: Mineralization activity, degradation pathways and boosted charge separation mechanism. Chem. Eng. J. 2021, 415, 128991. [Google Scholar] [CrossRef]
  7. Li, L.X.; Sun, X.F.; Xian, T.; Gao, H.J.; Wang, S.F.; Yi, Z.; Wu, X.W.; Yang, H. Template-free synthesis of Bi2O2CO3 hierarchical nanotubes self-assembled from ordered nanoplates for promising photocatalytic applications. Phys. Chem. Chem. Phys. 2022, 24, 8279–8295. [Google Scholar] [CrossRef]
  8. Sall, M.L.; Diaw, A.K.D.; Gningue-Sall, D.; Aaron, S.E.; Aaron, J.J. Toxic heavy metals: Impact on the environment and human health, and treatment with conducting organic polymers, a review. Environ. Sci. Pollut. Res. 2020, 27, 29927–29942. [Google Scholar] [CrossRef]
  9. Zhou, Q.Q.; Yang, N.; Li, Y.Z.; Ren, B.; Ding, X.H.; Bian, H.L.; Yao, X. Total concentrations and sources of heavy metal pollution in global river and lake water bodies from 1972 to 2017. Glob. Ecol. Conserv. 2020, 22, e00925. [Google Scholar] [CrossRef]
  10. Alhalili, Z. Metal Oxides Nanoparticles: General Structural Description, Chemical, Physical, and Biological Synthesis Methods, Role in Pesticides and Heavy Metal Removal through Wastewater Treatment. Molecules 2023, 28, 3086. [Google Scholar] [CrossRef]
  11. Li, Y.X.; Han, Y.C.; Wang, C.C. Fabrication strategies and Cr(VI) elimination activities of the MOF-derivatives and their composites. Chem. Eng. J. 2021, 405, 126648. [Google Scholar] [CrossRef]
  12. Xu, J.J.; Gu, H.Y.; Chen, M.D.; Li, X.P.; Zhao, H.W.; Yang, H.B. Dual Z-scheme Bi3TaO7/Bi2S3/SnS2 photocatalyst with high performance for Cr(VI) reduction and TC degradation under visible light irradiation. Rare Met. 2022, 41, 2417–2428. [Google Scholar] [CrossRef]
  13. Komijani, M.; Shamabadi, N.S.; Shahin, K.; Eghbalpour, F.; Tahsili, M.R.; Bahram, M. Heavy metal pollution promotes antibiotic resistance potential in the aquatic environment. Environ. Pollut. 2021, 274, 116569. [Google Scholar] [CrossRef]
  14. Kang, C.H.; So, J.S. Heavy metal and antibiotic resistance of ureolytic bacteria and their immobilization of heavy metals. Ecol. Eng. 2016, 97, 304–312. [Google Scholar] [CrossRef]
  15. Seiler, C.; Berendonk, T.U. Heavy metal driven co-selection of antibiotic resistance in soil and water bodies impacted by agriculture and aquaculture. Front. Microbiol. 2012, 3, 399. [Google Scholar] [CrossRef] [Green Version]
  16. Yang, Z.; Jia, S.Y.; Zhang, T.T.; Zhuo, N.; Dong, Y.Y.; Yang, W.B.; Wang, Y.P. How heavy metals impact on flocculation of combined pollution of heavy metals-antibiotics: A comparative study. Sep. Purif. Technol. 2015, 149, 398–406. [Google Scholar] [CrossRef]
  17. Wang, Y.X.; Rao, L.; Wang, P.F.; Shi, Z.Y.; Zhang, L.X. Photocatalytic activity of N-TiO2/O-doped N vacancy g-C3N4 and the intermediates toxicity evaluation under tetracycline hydrochloride and Cr(VI) coexistence environment. Appl. Catal. B. 2020, 262, 118308. [Google Scholar] [CrossRef]
  18. Li, B.S.; Lai, C.; Xu, P.A.; Zeng, G.M.; Huang, D.L.; Qin, L.; Yi, H.; Cheng, M.; Wang, L.L.; Huang, F.L.; et al. Facile synthesis of bismuth oxyhalogen-based Z-scheme photocatalyst for visible-light-driven pollutant removal: Kinetics, degradation pathways and mechanism. J. Clean. Prod. 2019, 225, 898–912. [Google Scholar] [CrossRef]
  19. He, H.; Jiang, J.; Luo, Z.; Li, D.; Shi, M.; Sun, H.; Chen, J.; Chen, C.; Deng, B.; Yu, C. Novel starfish-like inorganic/organic heterojunction for Cr(VI) photocatalytic reduction in neutral solution. Colloids Surf. A 2023, 667, 131357. [Google Scholar] [CrossRef]
  20. Ye, J.E.; Liu, J.H.; Huang, Z.J.; Wu, S.Y.; Dai, X.D.; Zhang, L.; Cui, L.H. Effect of reduced graphene oxide doping on photocatalytic reduction of Cr(VI) and photocatalytic oxidation of tetracycline by ZnAlTi layered double oxides under visible light. Chemosphere 2019, 227, 505–513. [Google Scholar] [CrossRef]
  21. Zhao, W.; Li, J.; Dai, B.L.; Cheng, Z.P.; Xu, J.M.; Ma, K.R.; Zhang, L.L.; Sheng, N.; Mao, G.X.; Wu, H.W.; et al. Simultaneous removal of tetracycline and Cr(VI) by a novel three-dimensional AgI/BiVO4 p-n junction photocatalyst and insight into the photocatalytic mechanism. Chem. Eng. J. 2019, 369, 716–725. [Google Scholar] [CrossRef]
  22. He, H.; Xiao, J.; Liu, Z.; Yang, B.; Wang, D.; Peng, X.; Zeng, L.; Li, Z.; Lei, L.; Qiu, M.; et al. Boosting the hydrogen evolution of layered double hydroxide by optimizing the electronic structure and accelerating the water dissociation kinetics. Chem. Eng. J. 2023, 453, 139751. [Google Scholar] [CrossRef]
  23. Yang, J.Y.; Liu, J.X.; Shi, F.; Ran, S.; Liu, S.H. Tm, Yb co-doped urchin-like CsxWO3 nanoclusters with dual functional properties: Transparent heat insulation performance and enhanced photocatalysis. Ceram. Int. 2021, 47, 8345–8356. [Google Scholar] [CrossRef]
  24. Aljuaid, A.; Almehmadi, M.; Alsaiari, A.A.; Allahyani, M.; Abdulaziz, O.; Alsharif, A.; Alsaiari, J.A.; Saih, M.; Alotaibi, R.T.; Khan, I. g-C3N4 Based Photocatalyst for the Efficient Photodegradation of Toxic Methyl Orange Dye: Recent Modifications and Future Perspectives. Molecules 2023, 28, 3199. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, W.H.; Mohamed, A.R.; Ong, W.J. Z-Scheme Photocatalytic Systems for Carbon Dioxide Reduction: Where Are We Now? Angew. Chem. Int. Ed. 2020, 59, 22894–22915. [Google Scholar] [CrossRef] [PubMed]
  26. Chen, D.J.; Cheng, Y.L.; Zhou, N.; Chen, P.; Wang, Y.P.; Li, K.; Huo, S.H.; Cheng, P.F.; Peng, P.; Zhang, R.C.; et al. Photocatalytic degradation of organic pollutants using TiO2-based photocatalysts: A review. J. Clean. Prod. 2020, 268, 121725. [Google Scholar] [CrossRef]
  27. Pitre, S.P.; Overman, L.E. Strategic Use of Visible-Light Photoredox Catalysis in Natural Product Synthesis. Chem. Rev. 2022, 122, 1717–1751. [Google Scholar] [CrossRef]
  28. Huang, R.D.; Wu, J.; Zhang, M.L.; Liu, B.Q.; Zheng, Z.Q.; Luo, D.X. Strategies to enhance photocatalytic activity of graphite carbon nitride-based photocatalysts. Mater. Des. 2021, 210, 110040. [Google Scholar] [CrossRef]
  29. Chakhtouna, H.; Benzeid, H.; Zari, N.; Qaiss, A.K.; Bouhfid, R. Recent progress on Ag/TiO2 photocatalysts: Photocatalytic and bactericidal behaviors. Environ. Sci. Pollut. Res. 2021, 28, 44638–44666. [Google Scholar] [CrossRef]
  30. Yang, X.R.; Chen, Z.; Zhao, W.; Liu, C.X.; Qian, X.X.; Zhang, M.; Wei, G.Y.; Khan, E.; Ng, Y.H.; Ok, Y.S. Recent advances in photodegradation of antibiotic residues in water. Chem. Eng. J. 2021, 405, 126806. [Google Scholar] [CrossRef]
  31. Mao, Z.Y.; Chen, J.J.; Yang, Y.F.; Wang, D.J.; Bie, L.J.; Fahlman, B.D. Novel g-C3N4/CoO Nanocomposites with Significantly Enhanced Visible-Light Photocatalytic Activity for H-2 Evolution. ACS Appl. Mater. Interfaces. 2017, 9, 12427–12435. [Google Scholar] [CrossRef] [PubMed]
  32. Duraisamy, E.; Das, H.T.; Sharma, A.S.; Elumalai, P. Supercapacitor and photocatalytic performances of hydrothermally-derived Co3O4/CoO@carbon nanocomposite. New J. Chem. 2018, 42, 6114–6124. [Google Scholar] [CrossRef]
  33. Xu, X.M.; Meng, L.J.; Luo, J.; Zhang, M.A.; Wang, Y.T.; Dai, Y.X.; Sun, C.; Wang, Z.Y.; Yang, S.G.; He, H.; et al. Self-assembled ultrathin CoO/Bi quantum dots/defective Bi2MoO6 hollow Z-scheme heterojunction for visible light-driven degradation of diazinon in water matrix: Intermediate toxicity and photocatalytic mechanism. Appl. Catal. B 2021, 293, 120231. [Google Scholar] [CrossRef]
  34. Liu, B.; Ma, L.; Ning, L.C.; Zhang, C.J.; Han, G.P.; Pei, C.J.; Zhao, H.; Liu, S.Z.; Yang, H.Q. Charge Separation between Polar {111} Surfaces of CoO Octahedrons and Their Enhanced Visible-Light Photocatalytic Activity. ACS Appl. Mater. Interfaces 2015, 7, 6109–6117. [Google Scholar] [CrossRef] [PubMed]
  35. Guan, H.; Wang, X.; Li, H.Q.; Zhi, C.Y.; Zhai, T.Y.; Bando, Y.; Golberg, D. CoO octahedral nanocages for high-performance lithium ion batteries. Chem. Commun. 2012, 48, 4878–4880. [Google Scholar] [CrossRef] [PubMed]
  36. Li, Y.; Zhang, J.; Liu, Z.W.; Liu, M.M.; Lin, H.J.; Che, R.C. Morphology-dominant microwave absorption enhancement and electron tomography characterization of CoO self-assembly 3D nano-flowers. J. Mater. Chem. C 2014, 2, 5216–5222. [Google Scholar] [CrossRef]
  37. Yang, H.M.; Ouyang, J.; Tang, A.D. Single step synthesis of high-purity CoO nanocrystals. J. Phys. Chem. B 2007, 111, 8006–8013. [Google Scholar] [CrossRef]
  38. Chen, X.Y.; Sun, B.J.; Chu, J.Y.; Han, Z.; Wang, Y.; Du, Y.N.; Han, X.J.; Xu, P. Oxygen Vacancy-Induced Construction of CoO/h-TiO2 Z-Scheme Heterostructures for Enhanced Photocatalytic Hydrogen Evolution. ACS Appl. Mater. Interfaces. 2022, 14, 28945–28955. [Google Scholar] [CrossRef]
  39. Wang, Y.J.; Guo, S.H.; Xin, X.; Zhang, Y.Z.; Wang, B.L.; Tang, S.W.; Li, X.H. Effective interface contact on the hierarchical 1D/2D CoO/NiCo-LDH heterojunction for boosting photocatalytic hydrogen evolution. Appl. Surf. Sci. 2021, 549, 149108. [Google Scholar] [CrossRef]
  40. Afsharpour, M.; Elyasi, M.; Javadian, H. A Novel N-Doped Nanoporous Bio-Graphene Synthesized from Pistacia lentiscus Gum and Its Nanocomposite with WO3 Nanoparticles: Visible-Light-Driven Photocatalytic Activity. Molecules 2021, 26, 6569. [Google Scholar] [CrossRef]
  41. Song, B.; Wang, T.T.; Sun, H.G.; Shao, Q.; Zhao, J.K.; Song, K.K.; Hao, L.H.; Wang, L.; Guo, Z.H. Two-step hydrothermally synthesized carbon nanodots/WO3 photocatalysts with enhanced photocatalytic performance. Dalton Trans. 2017, 46, 15769–15777. [Google Scholar] [CrossRef]
  42. Wang, G.M.; Ling, Y.C.; Wang, H.Y.; Yang, X.Y.; Wang, C.C.; Zhang, J.Z.; Li, Y. Hydrogen-treated WO3 nanoflakes show enhanced photostability. Energy Environ. Sci. 2012, 5, 6180–6187. [Google Scholar] [CrossRef]
  43. Fan, Y.S.; Xi, X.L.; Liu, Y.S.; Nie, Z.R.; Zhao, L.Y.; Zhang, Q.H. Regulation of morphology and visible light-driven photocatalysis of WO3 nanostructures by changing pH. Rare Met. 2021, 40, 1738–1745. [Google Scholar] [CrossRef]
  44. Cao, S.; Yu, J.G.; Wageh, S.; Al-Ghamdi, A.A.; Mousavi, M.; Ghasemi, J.B.; Xu, F.Y. H-2-production and electron-transfer mechanism of a noble-metal-free WO3@ZnIn2S4 S-scheme heterojunction photocatalyst. J. Mater. Chem. A 2022, 10, 17174–17184. [Google Scholar] [CrossRef]
  45. Ge, J.H.; Sun, Y.G.; Chen, W.W.; Song, F.G.; Xie, Y.; Zheng, Y.Y.; Rao, P.H. Z-scheme heterojunction based on NiWO4/WO3 microspheres with enhanced photocatalytic performance under visible light. Dalton Trans. 2021, 50, 13801–13814. [Google Scholar] [CrossRef] [PubMed]
  46. Alkanad, K.; Hezam, A.; Shekar, G.C.S.; Drmosh, Q.A.; Kala, A.L.A.; Al-Gunaid, M.Q.A.; Lokanath, N.K. Magnetic recyclable alpha-Fe2O3-Fe3O4/Co3O4-CoO nanocomposite with a dual Z-scheme charge transfer pathway for quick photo-Fenton degradation of organic pollutants. Catal. Sci. Technol. 2021, 11, 3084–3097. [Google Scholar] [CrossRef]
  47. Jin, J.; Yu, J.G.; Guo, D.P.; Cui, C.; Ho, W.K. A Hierarchical Z-Scheme CdS-WO3 Photocatalyst with Enhanced CO2 Reduction Activity. Small 2015, 11, 5262–5271. [Google Scholar] [CrossRef]
  48. Tang, M.L.; Ao, Y.H.; Wang, P.F.; Wang, C. All-solid-state Z-scheme WO3 nanorod/ZnIn2S4 composite photocatalysts for the effective degradation of nitenpyram under visible light irradiation. J. Hazard. Mater. 2020, 387, 121713. [Google Scholar] [CrossRef]
  49. Low, J.X.; Jiang, C.; Cheng, B.; Wageh, S.; Al-Ghamdi, A.A.; Yu, J.G. A Review of Direct Z-Scheme Photocatalysts. Small Methods 2017, 1, 1700080. [Google Scholar] [CrossRef]
  50. Liang, C.; Niu, C.G.; Guo, H.; Huang, D.W.; Wen, X.J.; Yang, S.F.; Zeng, G.M. Combination of efficient charge separation with the assistance of novel dual Z-scheme system: Self-assembly photocatalyst Ag@AgI/BiOI modified oxygen-doped carbon nitride nanosheet with enhanced photocatalytic performance. Catal. Sci. Technol. 2018, 8, 1161–1175. [Google Scholar] [CrossRef]
  51. Wang, X.H.; Wang, X.H.; Huang, J.F.; Li, S.X.; Meng, A.; Li, Z.J. Interfacial chemical bond and internal electric field modulated Z-scheme S-v-ZnIn2S4/MoSe2 photocatalyst for efficient hydrogen evolution. Nat. Commun. 2021, 12, 4112. [Google Scholar] [CrossRef]
  52. Fukumura, T.; Sambandan, E.; Yamashita, H. Synthesis and VOC degradation ability of a CeO2/WO3 thin-layer visible-light photocatalyst. Mater. Res. Bull. 2017, 94, 493–499. [Google Scholar] [CrossRef]
  53. Yang, Z.L.; Pu, H.T.; Yin, J.L. Preparation and electrochromic property of covalently bonded WO3/polyvinylimidazole core-shell microspheres. J. Colloid Interface Sci. 2005, 292, 108–112. [Google Scholar] [CrossRef] [PubMed]
  54. Wang, W.Z.; Zhang, G.L. Synthesis and optical properties of high-purity CoO nanowires prepared by an environmentally friendly molten salt route. J. Cryst. Growth 2009, 311, 4275–4280. [Google Scholar] [CrossRef]
  55. Zheng, J.H.; Zhang, L. Incorporation of CoO nanoparticles in 3D marigold flower-like hierarchical architecture MnCo2O4 for highly boosting solar light photo-oxidation and reduction ability. Appl. Catal. B. 2018, 237, 1–8. [Google Scholar] [CrossRef]
  56. Shi, W.L.; Guo, F.; Li, M.Y.; Shi, Y.; Shi, M.J.; Yan, C. Constructing 3D sub-micrometer CoO octahedrons packed with layered MoS2 shell for boosting photocatalytic overall water splitting activity. Appl. Surf. Sci. 2019, 473, 928–933. [Google Scholar] [CrossRef]
  57. Bai, J.X.; Shen, R.C.; Jiang, Z.M.; Zhang, P.; Li, Y.J.; Li, X. Integration of 2D layered CdS/WO3 S-scheme heterojunctions and metallic Ti3C2 MXene-based Ohmic junctions for effective photocatalytic H-2 generation. Chin. J. Catal. 2022, 43, 359–369. [Google Scholar] [CrossRef]
  58. Kumar, A.; Sharma, G.; Kumari, A.; Guo, C.S.; Naushad, M.; Vo, D.V.N.; Iqbal, J.; Stadler, F.J. Construction of dual Z-scheme g-C3N4/Bi4Ti3O12/Bi4O5I2 heterojunction for visible and solar powered coupled photocatalytic antibiotic degradation and hydrogen production: Boosting via I-/I-3(-) and Bi3+/Bi5+ redox mediators. Appl. Catal. B. 2021, 284, 119808. [Google Scholar] [CrossRef]
  59. Wei, X.Q.; Wang, X.; Pu, Y.; Liu, A.N.; Chen, C.; Zou, W.X.; Zheng, Y.L.; Huang, J.S.; Zhang, Y.; Yang, Y.C.; et al. Facile ball-milling synthesis of CeO2/g-C3N4 Z-scheme heterojunction for synergistic adsorption and photodegradation of methylene blue: Characteristics, kinetics, models, and mechanisms. Chem. Eng. J. 2021, 420, 127719. [Google Scholar] [CrossRef]
  60. Lv, Y.; Liu, Y.K.; Chen, C.M.; Wang, T.H.; Zhang, M. Octopus tentacles-like WO3/C@CoO as high property and long life-time electrocatalyst for hydrogen evolution reaction. Electrochim. Acta 2018, 281, 1–8. [Google Scholar] [CrossRef]
  61. Liu, J.L.; Yang, Q.; Liu, J.K.; Luo, H.A. Enhanced photoelectrochemical water oxidation of WO3/R-CoO and WO3/B-CoO photoanodes with a type II heterojunction. J. Mater. Sci. 2021, 56, 8079–8090. [Google Scholar] [CrossRef]
  62. Ding, W.; Ansari, N.; Yang, Y.H.; Bachagha, K. Superiorly sensitive and selective H-2 sensor based on p-n heterojunction of WO3-CoO nanohybrids and its sensing mechanism. Int. J. Hydrog. Energy 2021, 46, 28823–28837. [Google Scholar] [CrossRef]
  63. Li, X.R.; Du, J.N.; Liu, J.; Xu, X.Q.; Dai, Y.; Yu, Y.; Yuan, L.; Xie, J.H.; Zou, J.L. Microrods-evolved WO3 nanospheres with enriched oxygen-vacancies anchored on dodecahedronal CoO(Co2+)@carbon as durable catalysts for oxygen reduction/evolution reactions. Appl. Surf. Sci. 2022, 601, 154195. [Google Scholar] [CrossRef]
  64. Guo, F.; Shi, W.L.; Zhu, C.; Li, H.; Kang, Z.H. CoO and g-C3N4 complement each other for highly efficient overall water splitting under visible light. Appl. Catal. B. 2018, 226, 412–420. [Google Scholar] [CrossRef]
  65. Lu, Z.Y.; Yu, Z.H.; Dong, J.B.; Song, M.S.; Liu, Y.; Liu, X.L.; Ma, Z.F.; Su, H.; Yan, Y.S.; Huo, P.W. Facile microwave synthesis of a Z-scheme imprinted ZnFe2O4/Ag/PEDOT with the specific recognition ability towards improving photocatalytic activity and selectivity for tetracycline. Chem. Eng. J. 2018, 337, 228–241. [Google Scholar] [CrossRef]
  66. Shi, Y.X.; Li, L.L.; Xu, Z.; Sun, H.R.; Guo, F.; Shi, W.L. One-step simple green method to prepare carbon-doped graphitic carbon nitride nanosheets for boosting visible-light photocatalytic degradation of tetracycline. J. Chem. Technol. Biotechnol. 2021, 96, 3122–3133. [Google Scholar] [CrossRef]
  67. Wen, X.J.; Shen, C.H.; Fei, Z.H.; Niu, C.G.; Lu, Q.; Guo, J.; Lu, H.M. Fabrication of a zinc tungstate-based a p-n heterojunction photocatalysts towards refractory pollutants degradation under visible light irradiation. Colloids Surf. A Physicochem. Eng. Asp. 2019, 573, 137–145. [Google Scholar] [CrossRef]
  68. Guo, F.; Huang, X.L.; Chen, Z.H.; Cao, L.W.; Cheng, X.F.; Chen, L.Z.; Shi, W.L. Construction of Cu3P-ZnSnO3-g-C3N4 p-n-n heterojunction with multiple built-in electric fields for effectively boosting visible-light photocatalytic degradation of broad-spectrum antibiotics. Sep. Purif. Technol. 2021, 265, 118477. [Google Scholar] [CrossRef]
  69. Iqbal, A.; Saidu, U.; Adam, F.; Sreekantan, S.; Yahaya, N.; Ahmad, M.N.; Ramalingam, R.J.; Wilson, L.D. Floating ZnO QDs-Modified TiO2/LLDPE Hybrid Polymer Film for the Effective Photodegradation of Tetracycline under Fluorescent Light Irradiation: Synthesis and Characterisation. Molecules 2021, 26, 2509. [Google Scholar] [CrossRef]
  70. Zhang, G.X.; Tan, Y.; Sun, Z.M.; Zheng, S.L. Synthesis of BiOCl/TiO2 heterostructure composites and their enhanced photocatalytic activity. J. Environ. Chem. Eng. 2017, 5, 1196–1204. [Google Scholar] [CrossRef]
  71. Bao, S.; Liang, H.; Li, C.P.; Bai, J. The synthesis and enhanced photocatalytic activity of heterostructure BiOCl/TiO(2)nanofibers composite for tetracycline degradation in visible light. J. Dispers. Sci. Technol. 2021, 42, 2000–2013. [Google Scholar] [CrossRef]
  72. Yang, G.; Qiu, P.; Xiong, J.Y.; Zhu, X.T.; Cheng, G. Facilely anchoring Cu2O nanoparticles on mesoporous TiO2 nanorods for enhanced photocatalytic CO2 reduction through efficient charge transfer. Chin. Chem. Lett. 2022, 33, 3709–3712. [Google Scholar] [CrossRef]
  73. Ding, Y.Y.; Zhang, J.Y.; Yang, Y.; Long, L.Z.; Yang, L.; Yan, L.J.; Kong, W.J.; Liu, F.C.; Lv, F.; Liu, J. Fully-depleted dual P-N heterojunction with type-II band alignment and matched build-in electric field for high-efficient photocatalytic hydrogen production. Int. J. Hydrog. Energy 2021, 46, 36069–36079. [Google Scholar] [CrossRef]
  74. Bose, R.; Manna, G.; Jana, S.; Pradhan, N. Ag2S-AgInS2: P-n junction heteronanostructures with quasi type-II band alignment. Chem. Commun. 2014, 50, 3074–3077. [Google Scholar] [CrossRef] [PubMed]
  75. Li, J.T.; Liu, H.J.; Liu, Z.; Yang, D.Q.; Zhang, M.Z.; Gao, L.N.; Zhou, Y.H.; Lu, C.Y. Facile synthesis of Z-scheme NiO/alpha-MoO3 p-n heterojunction for improved photocatalytic activity towards degradation of methylene blue. Arab. J. Chem. 2022, 15, 103513. [Google Scholar] [CrossRef]
  76. Bavani, T.; Madhavan, J.; Preeyanghaa, M.; Neppolian, B.; Murugesan, S. Construction of direct Z-scheme g-C3N4/Bi2WO6 heterojunction photocatalyst with enhanced visible light activity towards the degradation of methylene blue. Environ. Sci. Pollut. Res. 2023, 30, 10179–10190. [Google Scholar] [CrossRef]
  77. Afroz, K.; Moniruddin, M.; Bakranov, N.; Kudaibergenov, S.; Nuraje, N. A heterojunction strategy to improve the visible light sensitive water splitting performance of photocatalytic materials. J. Mater. Chem. A 2018, 6, 21696–21718. [Google Scholar] [CrossRef]
  78. Zhao, D.M.; Wang, Y.Q.; Dong, C.L.; Huang, Y.C.; Chen, J.; Xue, F.; Shen, S.H.; Guo, L.J. Boron-doped nitrogen-deficient carbon nitride-based Z-scheme heterostructures for photocatalytic overall water splitting. Nat. Energy 2021, 6, 388–397. [Google Scholar] [CrossRef]
  79. Chen, X.J.; Wang, J.; Chai, Y.Q.; Zhang, Z.J.; Zhu, Y.F. Efficient Photocatalytic Overall Water Splitting Induced by the Giant Internal Electric Field of a g-C3N4/rGO/PDIP Z-Scheme Heterojunction. Adv. Mater. 2021, 33, 2007479. [Google Scholar] [CrossRef]
  80. Ghafoor, S.; Hussain, S.Z.; Waseem, S.; Arshad, S.N. Photo-reduction of heavy metal ions and photo-disinfection of pathogenic bacteria under simulated solar light using photosensitized TiO2 nanofibers. RSC Adv. 2018, 8, 20354–20362. [Google Scholar] [CrossRef] [Green Version]
  81. He, H.B.; Luo, Z.Z.; Tang, Z.Y.; Yu, C.L. Controllable construction of ZnWO4 nanostructure with enhanced performance for photosensitized Cr(VI) reduction. Appl. Surf. Sci. 2019, 490, 460–468. [Google Scholar] [CrossRef]
  82. Shi, W.; Guo, F.; Wang, H.; Han, M.; Li, H.; Yuan, S.; Huang, H.; Liu, Y.; Kang, Z. Carbon dots decorated the exposing high-reactive (111) facets CoO octahedrons with enhanced photocatalytic activity and stability for tetracycline degradation under visible light irradiation. Appl. Catal. B Environ. 2017, 219, 36–44. [Google Scholar] [CrossRef]
  83. Wang, S.; Zhao, L.; Huang, W.; Zhao, H.; Chen, J.; Cai, Q.; Jiang, X.; Lu, C.; Shi, W. Solvothermal synthesis of CoO/BiVO4 p-n heterojunction with micro-nano spherical structure for enhanced visible light photocatalytic activity towards degradation of tetracycline. Mater. Res. Bull. 2021, 135, 111161. [Google Scholar] [CrossRef]
  84. Pei, C.Y.; Chen, Y.G.; Wang, L.; Chen, W.; Huang, G.B. Step-scheme WO3/CdIn2S4 hybrid system with high visible light activity for tetracycline hydrochloride photodegradation. Appl. Surf. Sci. 2021, 535, 147682. [Google Scholar] [CrossRef]
  85. Lu, C.Y.; Yang, D.Q.; Wang, L.T.; Wen, S.J.; Cao, D.L.; Tu, C.Q.; Gao, L.N.; Li, Y.L.; Zhou, Y.H.; Huang, W. Facile construction of CoO/Bi2WO6 p-n heterojunction with following Z-Scheme pathways for simultaneous elimination of tetracycline and Cr(VI) under visible light irradiation. J. Alloys Compd. 2022, 904, 164046. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of CoO, WO3, and WO3/CoO heterojunctions.
Figure 1. XRD patterns of CoO, WO3, and WO3/CoO heterojunctions.
Molecules 28 04727 g001
Figure 2. TEM of (a) CoO, (b) WO3; and (ce) TEM and HR-TEM of 70% WO3/CoO.
Figure 2. TEM of (a) CoO, (b) WO3; and (ce) TEM and HR-TEM of 70% WO3/CoO.
Molecules 28 04727 g002
Figure 3. High-resolution spectra: (a) Co 2p, (b) W 4f, (c) C 1s, and (d) O 1s of the 70% WO3/CoO heterojunction.
Figure 3. High-resolution spectra: (a) Co 2p, (b) W 4f, (c) C 1s, and (d) O 1s of the 70% WO3/CoO heterojunction.
Molecules 28 04727 g003
Figure 4. (a) UV-Vis DRS of CoO, WO3, and WO3/CoO; (b) Tauc plot; and (c,d) Mott–Schottky of CoO and WO3.
Figure 4. (a) UV-Vis DRS of CoO, WO3, and WO3/CoO; (b) Tauc plot; and (c,d) Mott–Schottky of CoO and WO3.
Molecules 28 04727 g004
Figure 5. The photocatalytic activity of (a) TC, (d) Cr(VI), and (g) TC and Cr(VI) of the as-prepared samples; first-order reaction kinetics of photocatalytic degradation of (b) TC, (e) Cr(VI), and (h) TC and Cr(VI); the reaction rate constant of photocatalytic activity to (c) TC and (f) Cr(VI); and (i) five-cycle curve of 70% WO3/CoO sample for TC and Cr(VI) degradation.
Figure 5. The photocatalytic activity of (a) TC, (d) Cr(VI), and (g) TC and Cr(VI) of the as-prepared samples; first-order reaction kinetics of photocatalytic degradation of (b) TC, (e) Cr(VI), and (h) TC and Cr(VI); the reaction rate constant of photocatalytic activity to (c) TC and (f) Cr(VI); and (i) five-cycle curve of 70% WO3/CoO sample for TC and Cr(VI) degradation.
Molecules 28 04727 g005
Figure 6. (a) Photocurrent behavior, (b) electrochemical resistance, (c) PL spectrum, and (d) time-resolved luminescence decline of CoO, WO3, and 70% WO3/CoO heterojunctions.
Figure 6. (a) Photocurrent behavior, (b) electrochemical resistance, (c) PL spectrum, and (d) time-resolved luminescence decline of CoO, WO3, and 70% WO3/CoO heterojunctions.
Molecules 28 04727 g006
Figure 7. (a) Active species trapping experiments for visible photocatalytic degradation of TC by FSCN, and (b) ESR spectra of FSCN under visible light and dark conditions.
Figure 7. (a) Active species trapping experiments for visible photocatalytic degradation of TC by FSCN, and (b) ESR spectra of FSCN under visible light and dark conditions.
Molecules 28 04727 g007
Figure 8. Potential pathways for TC degradation by 70% WO3/CoO heterojunctions.
Figure 8. Potential pathways for TC degradation by 70% WO3/CoO heterojunctions.
Molecules 28 04727 g008
Figure 9. Potential photocatalytic mechanism for TC and Cr(VI) cleanup over WO3/CoO p-n heterojunction.
Figure 9. Potential photocatalytic mechanism for TC and Cr(VI) cleanup over WO3/CoO p-n heterojunction.
Molecules 28 04727 g009
Figure 10. Schematic diagram of synthesis for as-prepared photocatalysts.
Figure 10. Schematic diagram of synthesis for as-prepared photocatalysts.
Molecules 28 04727 g010
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lu, C.; Cao, D.; Zhang, H.; Gao, L.; Shi, W.; Guo, F.; Zhou, Y.; Liu, J. Boosted Tetracycline and Cr(VI) Simultaneous Cleanup over Z-Scheme WO3/CoO p-n Heterojunction with 0D/3D Structure under Visible Light. Molecules 2023, 28, 4727. https://doi.org/10.3390/molecules28124727

AMA Style

Lu C, Cao D, Zhang H, Gao L, Shi W, Guo F, Zhou Y, Liu J. Boosted Tetracycline and Cr(VI) Simultaneous Cleanup over Z-Scheme WO3/CoO p-n Heterojunction with 0D/3D Structure under Visible Light. Molecules. 2023; 28(12):4727. https://doi.org/10.3390/molecules28124727

Chicago/Turabian Style

Lu, Changyu, Delu Cao, Hefan Zhang, Luning Gao, Weilong Shi, Feng Guo, Yahong Zhou, and Jiahao Liu. 2023. "Boosted Tetracycline and Cr(VI) Simultaneous Cleanup over Z-Scheme WO3/CoO p-n Heterojunction with 0D/3D Structure under Visible Light" Molecules 28, no. 12: 4727. https://doi.org/10.3390/molecules28124727

Article Metrics

Back to TopTop