Next Article in Journal
Marine Algae-Derived Porous Carbons as Robust Electrocatalysts for ORR
Previous Article in Journal
Surface Display—An Alternative to Classic Enzyme Immobilization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In-Situ Construction of 2D/2D ZnIn2S4/BiOCl Heterostructure with Enhanced Photocatalytic Activity for N2 Fixation and Phenol Degradation

1
College of Chemistry and Chemical Engineering, Shaanxi Key Laboratory of Chemical Reaction Engineering, Yan’an University, Yan’an 716000, China
2
State Key Laboratory of Organic-Inorganic Composites Beijing Key Laboratory of Electrochemical Process and Technology for Materials, Beijing University of Chemical Technology, Beijing 100029, China
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(9), 729; https://doi.org/10.3390/catal9090729
Submission received: 9 August 2019 / Revised: 24 August 2019 / Accepted: 27 August 2019 / Published: 28 August 2019

Abstract

:
Photocatalytic nitrogen fixation and pollutants degradation at ambient temperature and pressure as a clean and efficient means represents a sustainable direction. In this paper, novel two-dimensional (2D) ZnIn2S4/2D BiOCl (ZIS/BOC) nanosheets heterostructures were successfully synthesized by in-situ solvothermal techniques. XPS and TEM results showed the formation of heterostructure. The photocatalytic properties of these materials were further studied by degradating phenol and nitrogen fixation under visible light irradiation. In particular, heterostructure obtained with a mass ratio of 0.5 wt % of ZnIn2S4 had a good application prospect and can achieve 77.4% phenol degradation rate within 6 h and a rate of ammonium production of 14.6 µmol∙gcat−1·h−1. The reaction rate of 0.5 wt % ZIS/BOC heterostructure exhibited 8.6 and 2.6 times relative to BiOCl for the degradation of phenol and nitrogen fixation, respectively. The enhanced photocatalytic activity was attributed to the collaboration of low recombination rate of photo-generated carriers and enhanced visible light absorption. Based on capture experiments and electron spin resonance (ESR) techniques, the main active species that may appear in photocatalytic systems were explored, and the corresponding photocatalytic mechanism was also proposed appropriately. This strategy will open up new ideas for the design and manufacture of more energy-efficient materials with high photocatalytic activity.

Graphical Abstract

1. Introduction

Rapid economic development and improvement of technology have an important impact on human life and quality of life. Among them, phenolic compounds have brought great convenience to human beings since they were found. However, wastewater generated in the production process is a very difficult problem and also has an indelible impact on the environment or even human health. Among phenolic-containing wastewater, phenol has the most severe toxicity. Therefore, efficient treatment of phenol compounds is of great significance to national economic and social development [1,2,3,4,5]. Moreover, nitrogen is widely found in nature and has an important impact on human life and quality of life. Particularly, efficient ammonia synthesis by chemical reaction involving nitrogen is of great significance to national economic and social development. The traditional Haber-Bosch process requires a lot of energy every year, so the search for a new high-efficiency, energy-saving, and environmentally-friendly nitrogen fixation method has attracted widespread public attention [6,7,8,9,10]. In recent decades, scientists have been working hard to find effective means to treat phenol compounds and achieve nitrogen fixation, and there has been little success [11,12,13,14,15,16,17,18,19,20,21].
In nature, green plants use solar light energy to assimilate carbon dioxide and water they absorb into organic matter. Photocatalytic as a kind of advanced oxidation process (AOPs) has attracted more and more attention recently because of its high efficiency and cleanness. Furthermore, photocatalytic degradation of pollutants and nitrogen fixation have been proved in various photocatalysts [22,23,24,25,26,27,28,29]. BiOCl, as one of the representatives of ternary oxides, has a special anisotropic layered structure with the band gap width is about 3.2 eV, and has good photocatalytic performance under ultraviolet irradiation, which is a bismuth-based compound with low cost, easy preparation, stable structure, and no toxicity, and has recently attracted widespread attention [30,31,32,33]. However, BiOCl itself cannot respond to visible light. Expanding its visible light response activity through effective modification methods has become the key to promote its industrial application. In order to solve these problems, a series of strategies have been developed to improve the performance of BiOCl [30,31,32,33,34,35,36]. Polymetallic sulfide semiconductor photocatalysis materials have been extensively studied due to their good light absorption performance and adjustable band gap [37,38,39]. Among many reported photocatalytic materials, ZnIn2S4 is a direct band gap with band energy of 2.4–2.7 eV [40,41], has shown good visible light photocatalytic performance due to its unique structure and physical and chemical properties, and has overcome the disadvantage of easy corrosion of traditional sulfides. Unfortunately, the single ZnIn2S4 has low transmission and separation efficiency of photogenerated carriers, and the electron (e)/hole (h+) pair is easy to combine, thus greatly reducing its photocatalytic performance [42,43,44]. In view of the above factors, many researches have been carried out to try to solve the above shortcomings. Furthermore, the combination of ZnIn2S4 with other semiconductors to improve the photocatalytic performance of ZnIn2S4 has been verified [42,43,44,45,46,47,48,49].
In this work, 2D/2D ZIS/BOC heterostructures were fabricated by in-situ solvothermal methods, and under the irradiation of visible light, photocatalytic reaction was used to degrade phenol and fix nitrogen. The activity results show that the prepared ZIS/BOC heterostructures have high photocatalytic degradation of phenol and nitrogen fixation activity. The designed interface structure of ZIS/BOC heterostructures was conducive to the separation and transfer of e and h+, and then enhanced the photocatalytic activity. Finally, according to capture experiments and ESR techniques, the corresponding photocatalytic mechanism was proposed.

2. Results and Discussion

2.1. Characterization of Photocatalysts

The crystal structure of the prepared sample was investigated by XRD, as shown in Figure 1. The nine diffraction peaks at 2θ angles of 11.98°, 24.10°, 25.86°, 32.50°, 33.45°, 36.54°, 43.56°, 49.70°, and 58.60° correspond to (001), (002), (101), (110), (102), (003), (103), (113), and (212) crystal planes of BiOCl (JCPDS No.06-0249) [32], while the peaks at 2θ angles of 27.69°, 34.60°, 47.17°, respectively come from (102), (016), and (110) crystal planes of ZnIn2S4 (JCPDS No.65-2023) [44]. The 0.5 wt % ZIS/BOC heterostructure presented nearly the same XRD patterns, indicating that the coupling of ZnIn2S4 nanosheets showed no effect on the crystal structure of BiOCl nanosheets. However, the characteristic peaks of ZnIn2S4 were not observed in the composite sample, which may be due to its low relative strength or content.
Figure 2a presented the SEM image of BiOCl, where it can be seen that BiOCl is a nanosheet structure with a thickness of about 20 nm. From the inset plot of Figure 2b, it can be seen that pure ZnIn2S4 are nanoparticles with a size of about 10 nm. The morphology of 0.5 wt % ZIS/BOC heterostructure exhibits a novel 2D/2D heterostructure (Figure 2b), which may be due to the directional growth of ZnIn2S4 on the surface of 2D BiOCl structure. Figure 2c,d presented the element distribution diagram (EDS) of 0.5 wt % ZIS/BOC heterostructure. Moreover, the composite catalytic material contains six elements of Bi, Cl, O, S, In, and Zn can be seen from Figure 2e–j, the distribution of each element in the 0.5 wt % ZIS/BOC heterostructure is relatively uniform. Figure 3 exhibited the TEM electron micrographs of BiOCl and 0.5 wt % ZIS/BOC heterostructure. As can be seen from Figure 3a,b, the BiOCl obtained has a nanosheet structure. Figure 3c,d are the TEM photographs of 0.5 wt % ZIS/BOC heterostructure. Figure 3c clearly showed a 2D/2D structure of the two compounds of BiOCl and ZnIn2S4. From Figure 3d, it can be seen that the crystal plane spacing of 0.275 nm belongs to (110) crystal plane of BiOCl, and the crystal plane spacing of 0.321 nm belongs to (102) crystal plane of ZnIn2S4, indicating that BiOCl and ZnIn2S4 form the 2D/2D heterostructure.
XPS was used to further confirm the presence of all elements in the heterostructure. Figure 4 shows the XPS spectroscopy of 0.5 wt % ZIS/BOC heterostructure, pure BiOCl, and ZnIn2S4. Further analysis of Figure 4a by XPS spectroscopy revealed that the sample 0.5 wt % ZIS/BOC heterostructure contained six elements of Cl, O, Bi, Zn, In, and S. Pure BiOCl contained three elements of Cl, O, and Bi, while ZnIn2S4 contained three elements of Zn, In, and S. Figure 4b–g are high resolution XPS spectra corresponding to Bi 4f, O 1s, Cl 2p, Zn, 2p in 3d and S 2p, respectively. Figure 4b showed that the signal peaks with binding energies of 158.8 and 164.1 eV correspond to Bi 4f7/2 and Bi 4f5/2 in the 0.5 wt % ZIS/BOC composite, indicating the presence of Bi3+ [50]. Figure 4c showed that the binding energy located at 530.4 and 529.5 eV are lattice oxygen and surface hydroxyl oxygen, respectively [34,51]. Figure 4d showed that the characteristic peaks with binding energies of 199.1 and 197.5 eV correspond to Cl 2p1/2 and Cl 2p3/2 [50]. It is worth noting that the peak positions of Bi 4f, O 1s, and Cl 2p in the 0.5 wt % ZIS/BOC heterostructure have a certain shift with respect to BiOCl. The shift might result from a partial e transfer from ZnIn2S4 to BiOCl. Similarly, in Figure 4e–g, high resolution peaks of Zn 2p were not observed, and In 3d at 441.4 and 451.8 eV, respectively, indicates that it exists in the In3+ valence state. However, the S 2p may be because the location of binding energy is similar to that of the Bi 4f or the content is too low and not observed [43,52]. The change of banding energy of elements may be ascribed to the formation of chemical band between BiOCl and ZnIn2S4. Based on the XPS results and combination of literature report [53,54], owing to the strong affinity between Bi3+ and S2−, Bi-S bond at the interface between BiOCl and ZnIn2S4 may be formed during in-situ solvothermal process.

2.2. Optical Properties and Photocatalytic Performances of The Samples

Figure 5 are the UV-visible diffuse reflectance spectrum of the samples. As can be seen from Figure 5, BiOCl has the basic absorption edge of about 362 nm and does not respond to visible light. However, pure ZnIn2S4 has strong absorption in the visible light region with absorption wavelength of 566 nm. After assembling ZnIn2S4 nanosheets onto the surface of BiOCl nanosheets, the light response range of BiOCl was widened from 362 to 505 nm. The enhanced absorption in visible light region indicates that ZIS/BOC heterostructures can generate more photogenerated carriers compared to pure BiOCl and have potential capability for photocatalytic pollutants degradation and nitrogen fixation under visible light.
In order to evaluate the catalytic performance of samples, we have investigated efficacy of as-obtained BiOCl and 0.5 wt % ZIS/BOC for photocatalytic N2 fixation. The photocatalytic experiments are performed in N2-saturated water a under the irradiation of 300 W Xe lamp in ultrapure water at ambient conditions without any sacrificial agent. Figure 6 showed the performance of photocatalytic nitrogen fixation over as-prepared catalyst under visible light irradiation. As determined by the Nessler’s reagent method [55], it revealed that it can hardly produce NH4+ in the presence of only catalyst or pumping N2 alone. While the concentration of NH4+ increased linearly with the extension of the illumination time (Figure 6a). Therein, 0.5 wt % ZIS/BOC heterostructure achieved a rate of ammonium production of 14.6 µmol gcat−1·h−1, which is 2.6 folds higher than that of pure BiOCl (5.7 µmol gcat−1·h−1). Furthermore, neither N2H4 nor H2 has been detected during the photocatalytic process, implying the excellent selectivity for NH3. It is interesting that the rate of ammonium production is low in the first half hour for BiOCl, while 0.5 wt %ZIS/BOC exhibited the same NH4+ production rate during the visible light irradiation, which may be ascribed to two aspects. First, the photocatalytic N2 fixation is a multistep reaction. The activation of N2 may mainly occur in the first half hour by transferring the photoexcited e from BiOCl to chemisorbed N2, dissociating N≡N bond and generating N2H* intermediates with coupled protons [55]. Second, 0.5 wt % ZIS/BOC heterostructure may exhibit more efficient carriers transfer at the photocatalyst/solution interface than that of BiOCl, achieving faster react rate from intermediates to the final product (NH4+). Also, the NH4+ production rates are about 14.6, 10, 8.5 and 5.7 µmol gcat−1·h−1, respectively, on those catalysts in an order of 0.5 wt % ZIS/BOC > 0.8 wt % ZIS/BOC > 0.3 wt % ZIS/BOC > BiOCl (Figure 6b). As can be seen from Figure 6c, the rate of ammonium production of the prepared 0.5 wt % ZIS/BOC decreased from 14.6% to 13.7% after 5 cycles, which can be attributed to deactivation of catalyst surface sites. These results clearly show that the synthesized ZIS/BOC heterostructure has higher stability.
To further investigate the photocatalytic performance of as-prepared samples, the photocatalytic degradation activity of phenol on different photocatalysts are performed as shown in Figure 7. A 300 W metal halide lamp with a wavelength of less than 420 nm cutoff was as a visible light source, and the photocatalytic activity was evaluated using phenol as a simulated pollutant. As can be seen from Figure 7a, the concentration of phenol gradually decreases with the increase of time. The analysis results showed that pure BiOCl only degraded 13.3% of phenol under visible light irradiation for 360 min, and the pure ZnIn2S4 degraded 31.5% of phenol, and 0.3 wt % ZIS-/BOC heterostructure, 0.5 wt % ZIS/BOC heterostructure and 0.8 wt % ZIS/BOC heterostructure degraded 55.4%, 77.4% and 57.4% of phenol, respectively, which is consistent with that of N2 reduction. This may be due to the fact that excess ZnIn2S4 occupies the active site and the light-shielding effect of ZnIn2S4 [56]. Figure 7b is the plot of the apparent rate constants for different catalysts. It can be clearly seen that the apparent rate constant for the degradation of phenol by 0.5 wt % ZIS/BOC heterostructure is 0.00415 min−1, which is 8.6, 2.3, 1.8, and 1.7 times than that of pure BiOCl, ZnIn2S4, 0.3 wt % ZIS/BOC heterostructure, and 0.8 wt % ZIS/BOC heterostructure, respectively, indicating that BiOCl nanosheets modified with ZnIn2S4 nanosheets and can improve photocatalytic activity. Moreover, the stability was confirmed by cycle experiments using 0.5 wt % ZIS/BOC heterostructure as a photocatalyst for the degradation of phenol. As shown in Figure 8, after 5 cycles, the significant change of photocatalytic degradation rate of phenol had not been observed. In addition to this, it can be concluded that from Figure 8 that the XRD pattern after the photocatalytic reaction was approximately similar to that of the 0.5 wt % ZIS/BOC heterostructure before the photocatalytic reaction, showing that 0.5 wt % ZIS/BOC heterostructure had good stability. The activity of photocatalytic degradation of phenol by using ZnIn2S4/BiOCl heterostructures as photocatalyst were compared with previous literature report [57,58,59,60,61,62,63], as shown in Table 1. Although the as-prepared ZnIn2S4/BiOCl heterostructures exhibited the promoted visible-light-driven photocatalytic activities than that of pristine ZnIn2S4 or BiOCl, its activity still needs to be further improved by other employing strategies.

2.3. Photocatalytic Mechanism

In order to further study the migration and separation efficiency of photo-generated carriers in ZIS/BOC heterostructure, fluorescence spectroscopy was performed on the prepared photocatalysts. Figure 9a showed the photoluminescence (PL) spectra of BiOCl and 0.5 wt % ZIS/BOC heterostructure at 280 nm excitation wavelength, and BiOCl has a strong fluorescence emission peak at about 391 nm. The PL intensity decreased obviously with the introduction of ZnIn2S4, indicating the effective separation of photoexcited e and h+. To reveal the generation and separation of photogenerated charges in the samples, the photoelectrochemical properties of the samples were investigated. As shown in Figure 9b, the fast photocurrent response of the sample after visible light irradiation was observed. Moreover, the photocurrent of 0.5 wt % ZIS/BOC heterostructure is higher compared with BiOCl. Figure 9c showed the impedance spectra of BiOCl and ZIS/BOC heterostructure, with an equivalent circuit model in the inset. The ohmic resistance (R1), corresponding to the intercept with the X-axis in high frequency, includes the intrinsic resistance of the photocatalyst and the electrolyte, as well as the contact resistance at the interface between the photocatalyst and the FTO glass. The semicircle in the middle frequency range corresponds to the charge transfer resistance (R2). The W and CPE represent the Warburg impedance and capacitance for the double layer between photocatalyst and electrolyte, respectively [64]. Comparing the Nyquist plots of BiOCl and ZIS/BOC (Figure 9c), it is found the 0.5 wt % ZIS/BOC electrode exhibits similar ohmic resistance (R1) due to the low content of ZIS in 0.5 wt % ZIS/BOC heterostructure, while 0.5 wt % ZIS/BOC heterostructure exhibits a much lower charge transfer resistance (R2) value (ca. 9116 Ω) than that of BiOCl (22337 Ω), implying the more efficient carriers transfer at the photocatalyst/solution interface, which is beneficial for the photocatalysis reaction [54]. Figure 10 is a schematic diagram of the effect of free radical scavenger on the photocatalytic activity of 0.5 wt % ZIS/BOC heterostructure. The three capture agents are benzoquinone (BQ, O2), ethylenediaminetetraacetic acid (EDTA, h+) and isopropanol (IPA,·OH), respectively. It can be clearly seen from Figure 10 that the addition of the three capture agents had a certain effect on the degradation rate of phenol compared to the absence of the capture agent. Among them, the addition of BQ minimized the degradation of phenol, which obviously inhibited the photocatalytic degradation rate of phenol. It is concluded that O2 was the main active species for phenol degradation. For the addition of EDTA and IPA, the activity was also inhibited, confirming that h+ and OH were also involved in the photocatalytic reaction.
The ESR capture experiment further confirmed the formation of active species. As shown in Figure 11, pure BiOCl showed substantially no signal of ·O2 and OH under visible light or in the dark, while 0.5 wt % ZIS/BOC heterostructure can produce significantly the typically 1:1:1:1 signal of O2, and the four ESR signals of 1:2:2:1 signal of OH under visible light irradiation. For ZIS/BOC heterostructure, the signals of the two active species are gradually increased compared to pure BiOCl, further indicating that active species ·O2 and OH were involved in the oxidative degradation of phenol. This is consistent with the captured experiments.
The bandgap of a semiconductor can be calculated by the following formula: (αhv)1/n = A(hν-Eg), where α, h, ν, a, and Eg respectively represent absorption coefficient, Planck constant, photon frequency, constant, and band-gap energy. Since BiOCl is an indirect transition semiconductor, it is calculated by substituting n = 2. Conversely, n = 0.5 for direct transition of ZnIn2S4. The band gap of BiOCl and ZnIn2S4 are 3.20 and 2.61 eV, respectively (Figure 12a,b). Meanwhile, BiOCl and ZnIn2S4 were determined to be 1.88 and 1.25 eV, respectively, according to XPS value results (Figure 12c,d). The calculated band position of the photocatalysts were shown in Table 2.
Figure 13 is a diagram showing the photocatalytic mechanism of the ZIS/BOC heterostructures. Appropriate conduction band and valence band positions can effectively promote the separation of photogenerated carriers when 2D BiOCl nanosheets and 2D ZnIn2S4 nanosheets are combined to form heterostructures. The conduction band position of ZnIn2S4 (−1.36 eV) is more negative than the conduction band position of BiOCl (−1.32 eV), so the conduction band e of ZnIn2S4 can transfer to the conduction band of BiOCl. In the process of photocatalytic degradation phenol, the conduction band position of ZnIn2S4 (−1.36 eV) is lower than O2/·O2 potential (−0.33 eV vs. NHE) so that O2 can obtain e to generate O2. However, since the valence band potential of ZnIn2S4 (1.25 eV) is lower than the redox potential of·OH/OH (1.99 eV) and the results of ESR and free radical trapping experiments both unexpectedly confirmed the existence of ·OH, which may be due to the partial conversion of O2 to OH, thereby increasing the photocatalytic activity of ZIS/BOC heterostructures. The mechanism is as follows (Equations (1)–(5)):
ZnIn2S4/BiOCl + hν → ZnIn2S4/BiOCl (e + h+ )
O2 + e→O2
O2 + e + 2H+→H2O2
H2O2 + O2→OH + OH + O2
H2O2 + e→OH + OH
As for the photocatalytic N2 fixation, the conduction band position of the ZnIn2S4 (−1.36 eV) is more negative than E0 (N2/NH3 = −0.28 eV vs. NHE), so the photogenerated e can further reduce the nitrogen that adsorbed on the surface of 0.5 wt % ZIS/BOC heterostructures to NH4+. In detail, the chemisorbed N2 molecule is activated by photogenerated e at the photocatalyst/solution interface [55], which is conducive to dissociating N≡N bond and generating N2H* intermediates with coupled protons (Equation (6)), and NH3* species are eventually generated (Equation (7)), which then exists in water. Finally, it was combined with H2O molecules to form NH4+ in succession for cycles (Equation (8)). The mechanism is as follows:
e + H+ + N2→N2H*
e + H+ + N2H*→NH3*
H+ + H2O + NH3*→NH4+

3. Experimental Section

3.1. Chemicals

All the reagents were of analytical grade and were used without any further purification.

3.2. Sample Preparation

Preparation of BiOCl nanosheets. BiOCl nanosheets were synthesized by an improved solvothermal method according to the previous report [65]. In a typical experiment, 0.9702 g (2 mmol) of Bi(NO3)3·5H2O and 0.1168 g (2 mmol) of NaCl were dissolved in 10 mL of ethylene glycol respectively, and magnetically stirred for 30 min to obtain clear and transparent solutions, which are denoted as A and B respectively. Subsequently, 0.1094 g of mannitol and 0.4 g of polyvinylpyrrolidone (PVP) were dissolved in 55 mL of water to form solution C. Firstly, solution A is added drop by drop into C, and solution B is slowly added to generate white precipitate after stirring for 5 min. The mixture was then transferred to a 100 mL Teflon-lined stainless-steel autoclave for hydrothermal reaction at 160 °C for 8 h. After the reaction was finished, the autoclave was naturally cooled to room temperature, and 15% NaOH aqueous solution was added to adjust pH to 13 after washing with water. The suspension was magnetically stirred for 3 h and washed with water to neutrality. The final samples were dried at 60 °C for 8 h, and then ground to obtain BiOCl powder.
Preparation of ZnIn2S4/BiOCl (ZIS/BOC) heterostructures. ZIS/BOC heterostructures were synthesized by an improved solvothermal method according to the previous report [66]. Firstly, 0.0036 g (0.0059 mmol) of Zn(NO3)2·6H2O, 0.0045 g (0.0118 mmol) of In(NO3)3·4.5H2O and 0.0018 g (0.0236 mmol) of Thioacetamide (TAA) were dissolved in 30 mL of ethylene glycol solution. After 20 min of magnetic stirring, 1 g of BiOCl was added to it. After 10 min of ultrasound, magnetic stirring was continued for 30 min. Then, the mixture was transferred to a 50 mL high pressure reactor lined with polytetrafluoroethylene for solvothermal reaction at 160 °C for 3 h under programmed temperature control, and the temperature was reduced to room-temperature. The yellow powder obtained after grinding is marked as 0.5 wt % ZnIn2S4/BiOCl. The amounts of Zn(NO3)2·6H2O, In(NO3)3·4.5H2O and TAA were changed to obtain 0.3 wt % and 0.8 wt % of composite catalyst ZnIn2S4/BiOCl respectively. The composite catalysts were labeled as 0.3 wt % ZIS/BOC, 0.5 wt % ZIS/BOC, and 0.8 wt %ZIS/BOC. Pure ZnIn2S4 was prepared by the same method without the addition of BiOCl.

3.3. Characterization

X-ray diffraction (XRD) was carried out on an XRD-7000 X-ray diffractometer (Shimadzu, Kyoto, Japan), using Cu Ka radiation (λ = 0.15418 nm). X-ray photoelectron spectroscopy (XPS) was recorded on a PHI-5400 X-ray photoelectron spectrometer (Physical Electronics PHI, Minnesota, America). Field emission scanning electron microscope (FESEM) images were recorded on a JSM-7610F scanning electron microscope (Japan electronics, Kyoto, Japan). Transmission electron microscopy (TEM) images were obtained on a JEM-2100 electron microscope (JEOL, Kyoto, Japan). The UV-vis diffuse reflectance spectra (UV-Vis-DRS) were measured with a UV-2550 UV-Vis spectrophotometer (Shimadzu, Kyoto, Japan). Photoluminescence (PL) spectra were measured on an F-4500 spectrophotometer (Physical Electronics PHI, Minnesota, America). Photocurrent density versus time (i–t) curves and electrochemical impedance spectroscopy (EIS) were measured on a CHI660E electrochemical workstation (Chenhua Instruments, Shanghai, China). The electron spin resonance (ESR) spectra were obtained on a JES-FA300 model spectrometer (ESR, Japan JEOL) under visible light irradiation (λ ≥ 420 nm).

3.4. Photocatalytic Activity Measurement

Photocatalytic degradation experiment. In the photocatalytic degradation experiment, phenol was used as a simulated pollutant to evaluate the photocatalytic activity of the samples. A suspension containing 0.2 g of photocatalyst is added to 200 mL of 10 mg·L−1 phenol solution, then placed in a photo reactor, and stirred in the dark until adsorption/desorption equilibrium is reached. A 300 W metal halogen lamp with a 420 nm cut-off filter was chosen as the visible light source. After turning on the light source, the supernatant is taken out according to a certain time interval. Finally, the absorbance of different solution concentrations was measured by using the 4-aminontipyrinolimetric method.
Photocatalytic Nitrogen Fixation. In nitrogen fixation experiment, a 300 W xenon lamp (with cut-off filter λ < 420 nm) was used as the light source, and a 500 mL reactor was equipped with circulating cooling water to ensure constant temperature. In general, 200 mg of photocatalyst is uniformly dispersed in 200 mL of ultrapure water, and high-purity N2 is pumped into the reactor for 30 min to saturate the suspension and remove the dissolved oxygen under dark conditions. Subsequently, the light source was turned on, 10 mL of supernatant was taken out every 30 min, and then centrifuged at 8500 rpm for 10 min to remove the photocatalyst. The concentration of ammonia was finally measured by Nessler’s reagent spectrophotometry method [55].

4. Conclusions

In conclusion, 2D BiOCl nanosheets were used as a substrate for the in-situ growth of the ZnIn2S4 nanosheets, and the carrier transport of the ZnIn2S4 is promoted based on the 2D/2D feature of the substrate. The photocatalytic phenol degradation and nitrogen fixation efficiency of ZIS/BOC heterostructures is significantly higher than that of BiOCl. It was determined that 0.5% ZIS/BOC heterostructure was the best ratio with excellent photocatalytic performance when phenol degradation was tested. The photocatalytic degradation efficiency of phenol was 77.4% and the ammonia production rate was 14.6 µmol∙gcat−1∙h−1 under visible light irradiation. The improved photocatalytic performance is attributed to the effective transfer and separation of photogenerated carriers at the interface and the enhanced light adsorb ability. The photocatalyst still maintained relatively high performance even after 5 cycles. The result of free radical trapping experiment and ESR analysis showed that O2 and·OH are the main active substances in visible light-driven catalytic system. In all, an appropriate mechanism is purposed.

Author Contributions

L.G. designed the experiments and wrote the paper; X.H. wrote the paper; K.Z., Y.Z. and Q.Z. performed the experiments and analyzed the data; D.W. and F.F. designed the experiments. All authors commented and approved the final manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (No. 21663030, 21666039) and the Open Project of State Key Laboratory of Organic–Inorganic Composites Beijing Key Laboratory, Beijing University of Chemical Technology Beijing (No. oic-201901009) and the Project of Science & Technology Office of Shaanxi Province (No. 2018TSCXL-NY-02-01, 2013SZS20-P01) and the Project of Yan’an Science and Technology Bureau (No. 2018KG-04) and Graduate Innovation Project of Yan’an University (YCX201988).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, W.; Zhou, J.; Zhang, Z.; Yu, J.; Cai, W. Different Surfactant-Assisted Hydrothermal Synthesis of Hierarchical Gamma-Al2O3 and Its Adsorption Performances for Parachlorophenol. Chem. Eng. J. 2013, 233, 168–175. [Google Scholar] [CrossRef]
  2. Sun, Y.; Zhou, J.; Cai, W.; Zhao, R.; Yuan, J. Hierarchically Porous NiAl-LDH Nanoparticles as Highly Efficient Adsorbent for p-Nitrophenol from Water. Appl. Surf. Sci. 2015, 349, 897–903. [Google Scholar] [CrossRef]
  3. Avetta, P.; Pensato, A.; Minella, M.; Malandrino, M.; Maurino, V.; Minero, C.; Hanna, K.; Vione, D. Activation of Persulfate by Irradiated Magnetite: Implications for the Degradation of Phenol under Heterogeneous Photo-Fenton-Like Conditions. Environ. Sci. Technol. 2015, 49, 104–1050. [Google Scholar] [CrossRef] [PubMed]
  4. Fu, F.; Shen, H.D.; Sun, X.; Xue, W.W.; Shoneyec, A.; Ma, J.N.; Luo, L.; Wang, D.J.; Wang, J.G.; Tang, J.W. Synergistic Effect of Surface Oxygen Vacancies and Interfacial Charge Transfer on Fe(III)/Bi2MoO6 for Efficient Photocatalysis. Appl. Catal. B 2019, 247, 150–162. [Google Scholar] [CrossRef]
  5. Ding, Z.; Lu, G.Q.; Greenfield, P.F. Role of the Crystallite Phase of TiO2 in Heterogeneous Photocatalysis for Phenol Oxidation in Water. J. Phys. Chem. B 2000, 104, 4815–4820. [Google Scholar] [CrossRef]
  6. Schrauzer, G.N.; Guth, T.D. Photocatalytic Reactions. 1. Photolysis of Water and Photoreduction of Nitrogen on Titanium Dioxide. J. Am. Chem. Soc. 1977, 99, 7189–7193. [Google Scholar] [CrossRef]
  7. Zhou, F.; Azofra, L.M.; Ali, M.; Kar, M.; Simonov, A.N.; McDonnell-Worth, C.; Sun, C.; Zhang, X.; MacFarlane, D.R. Electro-Synthesis of Ammonia from Nitrogen at Ambient Temperature and Pressure in Ionic Liquids. Energy Environ. Sci. 2017, 10, 2516–2520. [Google Scholar] [CrossRef]
  8. Rong, X.; Mao, Y.; Xu, J.; Zhang, X.; Zhang, L.; Zhou, X.; Qiu, F.; Wu, Z. Bi2Te3 Sheet Contributing to the Formation of Flower-Like BiOCl Composite and Its N2 Photofixation Ability Enhancement. Catal. Commun. 2018, 116, 16–19. [Google Scholar] [CrossRef]
  9. Feng, X.; Chen, H.; Jiang, F.; Wang, X. In-Situ Self-Sacrificial Fabrication of Lanthanide Hydroxycarbonates/Graphitic Carbon Nitride heterostructures: Nitrogen Photofixation under Simulated Solar Light Irradiation. Chem. Eng. J. 2018, 347, 849–859. [Google Scholar] [CrossRef]
  10. Chen, X.; Zhao, X.; Kong, Z.; Ong, W.J.; Li, N. Unravelling the Electrochemical Mechanisms for Nitrogen Fixation on Single Transition Metal Atoms Embedded in Defective Graphitic Carbon Nitride. J. Mater. Chem. A 2018, 6, 21941–21948. [Google Scholar] [CrossRef]
  11. Zhang, H.; Guo, L.H.; Wang, D.; Zhao, L.; Wan, B. Light-Induced Efficient Molecular Oxygen Activation on a Cu(II)–Grafted TiO2/Graphene Photocatalyst for Phenol Degradation. ACS Appl. Mater. Interfaces 2015, 7, 1816–1823. [Google Scholar] [CrossRef] [PubMed]
  12. Pradhan, G.K.; Padhi, D.K.; Parida, K.M. Fabrication of α-Fe2O3 Nanorod/RGO Composite: A Novel Hybrid Photocatalyst for Phenol Degradation. ACS Appl. Mater. Interfaces 2013, 5, 9101–9110. [Google Scholar] [CrossRef] [PubMed]
  13. Fu, F.; Shen, H.D.; Xue, W.W.; Zhen, Y.Z.; Soomro, R.A.; Yang, X.X.; Wang, D.J.; Xu, B.; Chi, R.A. Alkali-Assisted Synthesis of Direct Z-Scheme Based Bi2O3/Bi2MoO6 Photocatalyst for Highly Efficient Photocatalytic Degradation of Phenol and Hydrogen Evolution Reaction. J. Catal. 2019, 375, 399–409. [Google Scholar] [CrossRef]
  14. Saputra, E.; Muhammad, S.; Sun, H.Q.; Ang, H.M.; Tadé, M.O.; Wang, S.B. Manganese Oxides at Different Oxidation States for Heterogeneous Activation of Peroxymonosulfate for Phenol Degradation in Aqueous Solutions. Appl. Catal. B 2013, 142-143, 729–735. [Google Scholar] [CrossRef]
  15. Saputra, E.; Muhammad, S.; Sun, H.Q.; Ang, H.M.; Tadé, M.O.; Wang, S.B. Shape-Controlled Activation of Peroxymonosulfate by Single Crystal α-Mn2O3 for Catalytic Phenol Degradation in Aqueous Solution. Appl. Catal. B 2014, 154, 246–251. [Google Scholar] [CrossRef]
  16. Luo, Y.; Chen, G.; Ding, L.; Chen, X.; Ding, L.; Wang, H. Efficient Electrocatalytic N2 Fixation with MXene under ambient conditions. Joule 2019, 3, 1–11. [Google Scholar] [CrossRef]
  17. Hoffman, B.M.; Lukoyanov, D.; Yang, Z.Y.; Dean, D.R.; Seefeldt, L.C. Mechanism of Nitrogen Fixation by Nitrogenase: The Next Stage. Chem. Rev. 2014, 114, 4041–4062. [Google Scholar] [CrossRef] [PubMed]
  18. Honkala, K.; Hellman, A.; Remediakis, I.N.; Logadottir, A.; Carlsson, A.; Dahl, S.; Christensen, C.H.; Nørskov, J.K. Ammonia Synthesis from First-Principles Calculations. Science 2005, 307, 555–558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Tanabe, Y.; Nishibayashi, Y. Developing More Dustainable Processes for Ammonia Synthesis. Coord. Chem. Rev. 2013, 257, 2551–2564. [Google Scholar] [CrossRef]
  20. Sivasankar, C.; Baskaran, S.; Tamizmani, M.; Ramakrishna, K. Lessons Learned and Lessons to be Learned for Developing Homogeneous Transition Metal Complexes Catalyzed Reduction of N2 to Ammonia. J. Organomet. Chem. 2014, 752, 44–58. [Google Scholar] [CrossRef]
  21. Ma, H.; Shi, Z.; Li, Q.; Li, S. Preparation of Graphitic Carbon Nitride with Large Specifific Surface Area and Outstanding N2 Photofifixation Ability via a Dissolve-Regrowth Process. J. Phys. Chem. Solids 2016, 99, 51–58. [Google Scholar] [CrossRef]
  22. Qiu, P.; Xu, C.; Zhou, N.; Chen, H.; Jiang, F. Metal-Free Black Phosphorus Nanosheets-Decorated Graphitic Carbon Nitride Nanosheets with C-P Bonds for Excellent Photocatalytic Nitrogen Fixation. Appl. Catal. B 2018, 221, 27–35. [Google Scholar] [CrossRef]
  23. Liu, Y.; Kong, J.; Yuan, J.; Zhao, W.; Zhu, X.; Sun, C.; Xie, J. Enhanced Photocatalytic Activity over Flower-like Sphere Ag/Ag2CO3/BiVO4 Plasmonic Heterostructure Photocatalyst for Tetracycline Degradation. Chem. Eng. J. 2018, 331, 242–254. [Google Scholar] [CrossRef]
  24. Jiang, B.; Wang, Y.; Wang, J.; Tian, C.; Li, W.; Feng, Q.; Pan, Q.; Fu, H. In Situ Fabrication of Ag/Ag3PO4/Graphene Triple Heterostructure Visible-Light Photocatalyst Through Graphene-assisted Reduction Strategy. ChemCatChem 2013, 5, 1359–1367. [Google Scholar] [CrossRef]
  25. Ma, D.K.; Guan, M.L.; Liu, S.S.; Zhang, Y.Q.; Zhang, C.W.; He, Y.X.; Huang, S.M. Controlled Synthesis of Olive-shaped Bi2S3/BiVO4 Microspheres through a Limited Chemical Conversion Route and Enhanced Visible-Light-Responding Photocatalytic Activity. Dalton Trans. 2012, 41, 5581–5586. [Google Scholar] [CrossRef] [PubMed]
  26. Guo, Y.; Li, J.; Gao, Z.; Zhu, X.; Liu, Y.; Wei, Z.; Zhao, W.; Sun, C. A Simple and Effective Method for Fabricating Novel p–n Heterostructure Photocatalyst g-C3N4/Bi4Ti3O12 and Its Photocatalytic Performances. Appl. Catal. B 2016, 192, 57–71. [Google Scholar] [CrossRef]
  27. Miyama, H.; Fujii, N.; Nagae, Y. Heterogeneous Photocatalytic Synthesis of Ammonia from Water and Nitrogen. Chem. Phys. Lett. 1980, 74, 523–524. [Google Scholar] [CrossRef]
  28. Endoh, E.; Leland, J.K.; Bard, A.J. Heterogeneous Photoreduction of Nitrogen to Ammonia on Tungsten Oxide. J. Phys. Chem. 1986, 90, 6223–6226. [Google Scholar] [CrossRef]
  29. Hu, S.Z.; Chen, X.; Li, Q.; Li, F.Y.; Fan, Z.P.; Wang, H.; Wang, Y.J.; Zheng, B.H.; Wu, G. Fe3+ Doping Promoted N2 Photofixation Ability of Honeycombed Graphitic Carbon Nitride: The Experimental and Density Functional Theory Simulation Analysis. Appl. Catal. B 2017, 201, 58–69. [Google Scholar] [CrossRef]
  30. Jiang, J.; Zhao, K.; Xiao, X.Y.; Zhang, L.Z. Synthesis and Facet-Dependent Photoreactivity of BiOCl Single Crystalline Nanosheets. J. Am. Chem. Soc. 2012, 134, 4473–4476. [Google Scholar] [CrossRef]
  31. Jiang, R.R.; Wu, D.H.; Lu, G.H.; Yan, Z.H.; Liu, J.C.; Zhou, R.R.; Matthew, N. Fabrication of Fe3O4 Quantum Dots Modified BiOCl/BiVO4 p-n Heterostructure to Enhance Photocatalytic Activity for Removing Broad-Spectrum Antibiotics under Visible Light. J. Taiwan Inst. Chem. Eng. 2019, 96, 681–690. [Google Scholar] [CrossRef]
  32. Guan, M.L.; Xiao, C.; Zhang, J.; Fan, S.J.; An, R.; Cheng, Q.M.; Xie, J.F.; Zhou, M.; Ye, B.J.; Xie, Y. Vacancy Associates Promoting Solar-Driven Photocatalytic Activity of Ultrathin Bismuth Oxychloride Nanosheets. J. Am. Chem. Soc. 2013, 135, 10411–10417. [Google Scholar] [CrossRef] [PubMed]
  33. Deng, F.; Zhang, Q.; Yang, L.X.; Luo, X.B.; Wang, A.J.; Luo, S.L.; Dionysiou, D.D. Visible Light-Responsive Graphene-Functionalized Bi-Bridge Z-Scheme Black BiOCl/Bi2O3 Heterostructure with Oxygen Vacancy and Multiple Charge Transfer Channels for Efficient Photocatalytic Degradation of 2-Nitrophenol and Industrial Wastewater Treatment. Appl. Catal. B 2018, 238, 61–69. [Google Scholar] [CrossRef]
  34. Wu, X.; Zhang, K.; Zhang, G.; Yin, S. Facile Preparation of BiOX (X = Cl, Br, I) Nanoparticles and up-Conversion Phosphors/BiOBr Composites for Efficient Degradation of NO Gas: Oxygen Vacancy Effect and Near Infrared Light Responsive Mechanism. Chem. Eng. J. 2017, 325, 59–70. [Google Scholar] [CrossRef]
  35. Ye, L.; Liu, J.; Gong, C.; Tian, L.; Peng, T.; Zan, L. Two Different Roles of Metallic Ag on Ag/AgX/BiOX (X = Cl, Br) Visible Light Photocatalysts: Surface Plasmon Resonance and Z-Scheme Bridge. ACS Catal. 2012, 2, 1677–1683. [Google Scholar] [CrossRef]
  36. Myung, Y.; Wu, F.; Banerjee, S.; Park, J.; Banerjee, P. Electrical Conductivity of p-Type BiOCl Nanosheets. Chem. Commun. 2015, 51, 2629–2632. [Google Scholar] [CrossRef]
  37. Yao, L.; Wei, D.; Ni, Y.; Yan, D.; Hu, C. Surface Localization of CdZnS Quantum Dots onto 2D g-C3N4 Ultrathin Microribbons: Highly Efficient Visible Light-Induced H2-Generation. Nano Energy 2016, 26, 248–256. [Google Scholar] [CrossRef]
  38. Jing, L.; Xu, Y.; Xie, M.; Liu, J.; Deng, J.; Huang, L.; Xu, H.; Li, H. Three Dimensional Polyaniline/MgIn2S4 Nanoflower Photocatalysts Accelerated Interfacial Charge Transfer for the Photoreduction of Cr(VI), Photodegradation of Organic Pollution and Photocatalytic H2 Production. Chem. Eng. J. 2019, 360, 1601–1612. [Google Scholar] [CrossRef]
  39. Gou, X.L.; Cheng, F.Y.; Shi, Y.H.; Zhang, L.; Peng, S.J.; Chen, J.; Shen, P.W. Shape-Controlled Synthesis of Ternary Chalcogenide ZnIn2S4 and CuIn(S,Se)2 Nano-/Microstructures via Facile Solution Route. J. Am. Chem. Soc. 2006, 128, 7222–7229. [Google Scholar] [CrossRef]
  40. Zhou, M.; Liu, Z.H.; Song, Q.G.; Li, X.F.; Chen, B.W.; Liu, Z.F. Hybrid 0D/2D Edamame Shaped ZnIn2S4 Photoanode Modified by Co-Pi and Pt for Charge Management towards Efficient Photoelectrochemical Water Splitting. Appl. Catal. B 2019, 244, 188–196. [Google Scholar] [CrossRef]
  41. Fan, B.; Chen, Z.; Liu, Q.; Zhang, Z.; Fang, X. One-Pot Hydrothermal Synthesis of Ni-Doped ZnIn2S4 Nanostructured Film Photoelectrodes with Enhanced Photoelectrochemical Performance. Appl. Surf. Sci. 2016, 370, 252–259. [Google Scholar] [CrossRef]
  42. Guan, Z.; Pan, J.; Li, Q.; Li, G.; Yang, J. Boosting Visible-Light Photocatalytic Hydrogen Evolution with an Efficient CuInS2/ZnIn2S4 2D/2D Heterostructure. ACS Sustain. Chem. Eng. 2019, 7, 7736–7742. [Google Scholar] [CrossRef]
  43. Kale, S.B.; Kalubarme, R.S.; Mahadadalkar, M.A.; Jadhav, H.S.; Bhirud, A.P.; Ambekar, J.D.; Park, C.J.; Kale, B.B. Hierarchical 3D ZnIn2S4/Graphene Nano-Heterostructures: Their in Situ Fabrication with Dual Functionality in Solar Hydrogen Production and As Anodes for Lithium Ion Batteries. Phys. Chem. Chem. Phys. 2015, 17, 31850–31861. [Google Scholar] [CrossRef] [PubMed]
  44. Yang, C.; Li, Q.; Xia, Y.; Lv, K.; Li, M. Enhanced Visible-Light Photocatalytic CO2 Reduction Performance of Znln2S4 Microspheres by Using CeO2 as Cocatalyst. Appl. Surf. Sci. 2019, 464, 388–395. [Google Scholar] [CrossRef]
  45. Yang, G.; Chen, D.M.; Ding, H.; Feng, J.J.; Zhang, J.Z.; Zhu, Y.F.; Hamid, S.; Bahnemann, D.W. Well-Designed 3D ZnIn2S4 Nanosheets/TiO2 Nanobelts as Direct Z-Scheme photocatalysts for CO2 Photoreduction into Renewable Hydrocarbon Fuel with High Efficiency. Appl. Catal. B 2017, 219, 611–618. [Google Scholar] [CrossRef]
  46. Chai, B.; Peng, T.Y.; Zeng, P.; Zhang, X.H. Preparation of a MWCNTs/ZnIn2S4 Composite and Its Enhanced Photocatalytic Hydrogen Production under Visible-Light Irradiation. Dalton Trans. 2012, 41, 1179–1186. [Google Scholar] [CrossRef] [PubMed]
  47. Zhang, Z.Z.; Huang, L.; Zhang, J.J.; Wang, F.J.; Xie, Y.Y.; Shang, X.T.; Go, Y.Y.; Zhao, H.B.; Wang, X.X. In Situ Constructing Interfacial Contact MoS2/ZnIn2S4 Heterostructure for Enhancing Solar Photocatalytic Hydrogen Evolution. Appl. Catal. B 2018, 233, 112–119. [Google Scholar] [CrossRef]
  48. Yuan, Y.J.; Chen, D.Q.; Zhong, J.S.; Yang, L.X.; Wang, J.J.; Liu, M.J.; Tu, W.G.; Yu, Z.T.; Zou, Z.G. Interface Engineering of a Noble-metal-free 2D–2D MoS2/Cu-ZnIn2S4 Photocatalyst for Enhanced Photocatalytic H2 Production. J. Mater. Chem. A 2017, 5, 15771. [Google Scholar] [CrossRef]
  49. Sun, W.X.; Wei, N.; Cui, H.Z.; Lin, Y.; Wang, X.Z.; Tian, J.; Li, J.; Wen, J. 3D ZnIn2S4 Nanosheet/TiO2 Nanowire Arrays and Their Efficient Photocathodic Protection for 304 Stainless Steel. Appl. Surf. Sci. 2018, 434, 1030–1039. [Google Scholar] [CrossRef]
  50. Gao, X.Y.; Tang, G.B.; Peng, W.; Guo, Q.; Luo, Y.M. Surprise in the Phosphate Modification of BiOCl with Oxygen Vacancy: In Situ Construction of Hierarchical Z-Scheme BiOCl-OV-BiPO4 Photocatalyst for the Degradation of Carbamazepine. Chem. Eng. J. 2019, 360, 1320–1329. [Google Scholar] [CrossRef]
  51. Wang, D.J.; Guo, L.; Zhen, Y.Z.; Yue, L.L.; Xue, G.L.; Fu, F. AgBr Quantum Dots Decorated Mesoporous Bi2WO6 Architectures with Enhanced Photocatalytic Activities for Methylene Blue. J. Mater. Chem. A 2014, 2, 11716–11727. [Google Scholar] [CrossRef]
  52. Ruan, X.W.; Hu, H.; Che, H.N.; Jiang, E.H.; Zhang, X.X.; Liu, C.B.; Che, G.B. A Visible-Light-Driven Z-scheme CdS/Bi12GeO20 Heterostructure with Enhanced Photocatalytic Degradation of Various Organics and the Reduction of Aqueous Cr(VI). J. Colloid Interface Sci. 2019, 543, 317–327. [Google Scholar] [CrossRef] [PubMed]
  53. Bhoi, Y.P.; Mishra, B.G. Single Step Combustion Synthesis, Characterization and Photoctalytic Application of α-Fe2O3-Bi2S3 Heterojunction for Efficient and Selective Reduction of Structurally Diverse Nitroarenes. Chem. Eng. J. 2017, 316, 70–81. [Google Scholar] [CrossRef]
  54. Luo, S.; Ke, J.; Yuan, M.Q.; Zhang, Q.; Xie, P.; Deng, L.D.; Wang, S.B. CuInS2 Quantum Dots Embedded in Bi2WO6 Nanoflowers for Enhanced Visible Light Photocatalytic Removal of Contaminants. Appl. Catal. B 2017, 221, 215–222. [Google Scholar] [CrossRef]
  55. Zhang, N.; Jalil, A.; Wu, D.X.; Chen, S.M.; Liu, Y.F.; Gao, C.; Ye, W.; Qi, Z.M.; Ju, H.X.; Wang, C.M.; et al. Refining Defect States in W18O49 by Mo Doping: A Strategy for Tuning N2 Activation towards Solar-Driven Nitrogen Fixation. J. Am. Chem. Soc. 2018, 140, 9434–9443. [Google Scholar] [CrossRef] [PubMed]
  56. Ran, J.R.; Guo, W.W.; Wang, H.L.; Zhu, B.C.; Yu, J.G.; Qiao, S.Z. Metal-Free 2D/2D Phosphorene/g-C3N4 Van der Waals Heterojunction for Highly Enhanced Visible-Light Photocatalytic H2 Production. Adv. Mater. 2018, 30, 1800128. [Google Scholar] [CrossRef] [PubMed]
  57. Dewidar, H.; Nosier, S.A.; El-Shazly, A.H. Photocatalytic Degradation of Phenol Solution using Zinc Oxide/UV. J. Chem. Health Saf. 2018, 25, 2–11. [Google Scholar] [CrossRef]
  58. Chiou, C.H.; Juang, R.S. Photocatalytic Degradation of Phenol in Aqueous Solutions by Pr-Doped TiO2 Nanoparticles. J. Hazard. Mater. 2007, 149, 1–7. [Google Scholar] [CrossRef]
  59. Al-Hamdi, A.M.; Sillanpää, M.; Dutta, J. Photocatalytic Degradation of Phenol in Aqueous Solution by Rare Earth-Doped SnO2 Nanoparticles. J. Mater. Sci. 2014, 49, 5151–5159. [Google Scholar] [CrossRef]
  60. Sánchez-Rodríguez, D.; Medrano, M.G.M.; Remita, H.; Escobar-Barrios, V. Photocatalytic Properties of BiOCl-TiO2 Composites for Phenol Photodegradation. J. Environ. Chem. Eng. 2018, 6, 1601–1612. [Google Scholar] [CrossRef]
  61. Li, Q.B.; Zhao, X.; Yang, J.; Jia, C.J.; Jin, Z.; Fan, W.L. Exploring the Effects of Nanocrystal Facet Orientations in g-C3N4/BiOCl Heterostructures on Photocatalytic Performance. Nanoscale 2015, 7, 18971–18983. [Google Scholar] [CrossRef] [PubMed]
  62. Cai, L. Enhanced Visible Light Photocatalytic Activity of BiOCl by Compositing with g-C3N4. Mater. Res. Innov. 2015, 19, 392–396. [Google Scholar] [CrossRef]
  63. Liu, H.; Jin, Z.T.; Xu, Z.Z.; Zhang, Z.; Ao, D. Fabrication of ZnIn2S4-G-C3N4 Sheet-On-Sheet Nanocomposites for Efficient Visible-Light Photocatalytic H2 Evolution and Degradation of Organic Pollutants. RSC Adv. 2015, 5, 97951–97961. [Google Scholar] [CrossRef]
  64. Liu, H.; Jia, M.Q.; Sun, N.; Cao, B.; Chen, R.J.; Zhu, Q.Z.; Wu, F.; Qiao, N.; Xu, B. Nitrogen-Rich Mesoporous Carbon as Anode Material for High Performance Sodium-Ion Batteries. ACS Appl. Mater. Interfaces 2015, 7, 27124–27130. [Google Scholar] [CrossRef] [PubMed]
  65. Li, B.X.; Shao, L.Z.; Wang, R.S.; Dong, X.P.; Zhao, F.G.; Gao, P.; Li, Z.Q. Interfacial Synergism of Pd-Decorated BiOCl Ultrathin Nanosheets for the Selective Oxidation of Aromatic Alcohols. J. Mater. Chem. A 2018, 6, 6344–6355. [Google Scholar] [CrossRef]
  66. Lin, B.; Li, H.; An, H.; Hao, W.B.; Wei, J.J.; Dai, Y.Z.; Ma, C.S.; Yang, G.D. Preparation of 2D/2D g-C3N4 Nanosheet@ZnIn2S4 Nanoleaf Heterostructures with Well-Designed High-Speed Charge Transfer Nanochannels towards High-Efficiency Photocatalytic Hydrogen Evolution. Appl. Catal. B 2018, 220, 542–552. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction (XRD) patterns of (a) pure BiOCl, (b) 0.5 wt % ZIS/BOC, (c) pure ZnIn2S4.
Figure 1. X-ray diffraction (XRD) patterns of (a) pure BiOCl, (b) 0.5 wt % ZIS/BOC, (c) pure ZnIn2S4.
Catalysts 09 00729 g001
Figure 2. Scanning electron microscopy (SEM) patterns of (a) pure BiOCl, (b) 0.5 wt % ZIS/BOC and ZnIn2S4 (insert picture in Figure 2b), (cj) element distribution diagram (EDS) mapping of 0.5 wt % ZIS/BOC heterostructure.
Figure 2. Scanning electron microscopy (SEM) patterns of (a) pure BiOCl, (b) 0.5 wt % ZIS/BOC and ZnIn2S4 (insert picture in Figure 2b), (cj) element distribution diagram (EDS) mapping of 0.5 wt % ZIS/BOC heterostructure.
Catalysts 09 00729 g002
Figure 3. (a,b) Transmission electron microscopy (TEM) images of pure BiOCl and (c) 0.5 wt % ZIS/BOC heterostructure, and (d) HRTEM of 0.5 wt % ZIS/BOC heterostructure.
Figure 3. (a,b) Transmission electron microscopy (TEM) images of pure BiOCl and (c) 0.5 wt % ZIS/BOC heterostructure, and (d) HRTEM of 0.5 wt % ZIS/BOC heterostructure.
Catalysts 09 00729 g003
Figure 4. X-ray photoelectric spectroscopy (XPS) spectra of the BiOCl, ZnIn2S4 and 0.5 wt % ZIS/BOC heterostructure (a) survey, (b) Bi 4f, (c) O 1s and (d) Cl 2p of BiOCl and 0.5 wt % ZIS/BOC heterostructure, (e) Zn 2p, (f) In 3d, (g) S 2p of the 0.5 wt % ZIS/BOC heterostructure.
Figure 4. X-ray photoelectric spectroscopy (XPS) spectra of the BiOCl, ZnIn2S4 and 0.5 wt % ZIS/BOC heterostructure (a) survey, (b) Bi 4f, (c) O 1s and (d) Cl 2p of BiOCl and 0.5 wt % ZIS/BOC heterostructure, (e) Zn 2p, (f) In 3d, (g) S 2p of the 0.5 wt % ZIS/BOC heterostructure.
Catalysts 09 00729 g004
Figure 5. UV-vis diffuse reflectance spectra (UV-vis DRS spectra) of pure BiOCl, ZnIn2S4 and a series of ZIS/BOC heterostructures.
Figure 5. UV-vis diffuse reflectance spectra (UV-vis DRS spectra) of pure BiOCl, ZnIn2S4 and a series of ZIS/BOC heterostructures.
Catalysts 09 00729 g005
Figure 6. (a) Ammonia production amount of the BiOCl and 0.5 wt % ZIS/BOC heterostructure, (b) The rate of ammonium production on BiOCl and ZIS/BOC heterostructure. (c) Cycling times in the photocatalytic N2 fixation over 0.5 wt % ZIS/BOC heterostructure.
Figure 6. (a) Ammonia production amount of the BiOCl and 0.5 wt % ZIS/BOC heterostructure, (b) The rate of ammonium production on BiOCl and ZIS/BOC heterostructure. (c) Cycling times in the photocatalytic N2 fixation over 0.5 wt % ZIS/BOC heterostructure.
Catalysts 09 00729 g006
Figure 7. Photocatalytic degradation performance of BiOCl, ZnIn2S4 and the series of ZIS/BOC heterostructures for phenol (a), ln(C0/C) vs. reaction time (b), the apparent rate constants (c).
Figure 7. Photocatalytic degradation performance of BiOCl, ZnIn2S4 and the series of ZIS/BOC heterostructures for phenol (a), ln(C0/C) vs. reaction time (b), the apparent rate constants (c).
Catalysts 09 00729 g007
Figure 8. (a) Cycling times in the photocatalytic phenol degradation over 0.5 wt % ZIS/BOC heterostructure, (b) X-ray diffraction (XRD) pattern of 0.5 wt % ZIS/BOC heterostructure after and before used.
Figure 8. (a) Cycling times in the photocatalytic phenol degradation over 0.5 wt % ZIS/BOC heterostructure, (b) X-ray diffraction (XRD) pattern of 0.5 wt % ZIS/BOC heterostructure after and before used.
Catalysts 09 00729 g008
Figure 9. (a) The photoluminescence spectra, (b)Transient photocurrent response and (c) Electrochemical impedance spectroscopy (EIS) of the BiOCl and 0.5 wt % ZIS/BOC heterostructure.
Figure 9. (a) The photoluminescence spectra, (b)Transient photocurrent response and (c) Electrochemical impedance spectroscopy (EIS) of the BiOCl and 0.5 wt % ZIS/BOC heterostructure.
Catalysts 09 00729 g009
Figure 10. Photogenerated carrier trapping for degradation of phenol by 0.5 wt % ZIS/BOC heterostructure.
Figure 10. Photogenerated carrier trapping for degradation of phenol by 0.5 wt % ZIS/BOC heterostructure.
Catalysts 09 00729 g010
Figure 11. 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) spin-trapping electrochemical impedance spectroscopy (ESR) of·O2 and OH adducts resulting from visible light irradiation of BiOCl (a,c), and 0.5 wt % ZIS/BOC heterostructure (b,d).
Figure 11. 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) spin-trapping electrochemical impedance spectroscopy (ESR) of·O2 and OH adducts resulting from visible light irradiation of BiOCl (a,c), and 0.5 wt % ZIS/BOC heterostructure (b,d).
Catalysts 09 00729 g011aCatalysts 09 00729 g011b
Figure 12. (a,b) the corresponding Tauc curves of the BiOCl and ZnIn2S4, and (c,d) valence band spectra of the BiOCl and ZnIn2S4.
Figure 12. (a,b) the corresponding Tauc curves of the BiOCl and ZnIn2S4, and (c,d) valence band spectra of the BiOCl and ZnIn2S4.
Catalysts 09 00729 g012
Figure 13. The photocatalytic mechanism of the phenol degradation and nitrogen fixation over photocatalyst ZIS/BOC heterostructures under visible light irradiation.
Figure 13. The photocatalytic mechanism of the phenol degradation and nitrogen fixation over photocatalyst ZIS/BOC heterostructures under visible light irradiation.
Catalysts 09 00729 g013
Table 1. Comparison of photocatalytic performance for removal of phenol.
Table 1. Comparison of photocatalytic performance for removal of phenol.
SamplesLight SourceOrganic PollutantsRef.
Type (a)Power (W)Conc.(b)
(mg·L−1)
Catalyst
(g/L)
t (c)
(time/min)
η (d)
(%)
ZnOUV lamp15251.512078.2[57]
Pr-doped-TiO2high-pressure mercury lamp400501.012094.4[58]
La-doped-SnO2Mercury lamp8101.390100[59]
BiOCl/TiO2Xe lamp300501.036043[60]
g-C3N4/BiOClMetal halide lamp300300.648050[61]
g-C3N4/BiOClLED lamp5×24101.015094[62]
ZnIn2S4/g-C3N4Xe lamp500100.424072.3[63]
0.5% wt % ZIS/BOCMetal halide lamp300101.036077.4This work
(a) UV-lamp and mercury lamp were the UV-light sources, Xe lamp, LED lamp and metal halide lamp were the visible-light source; (b) concentration of pollutants(mg·L−1); (c) irradiation time (t/min); (d) the total removal efficiency of organic pollutant (%).
Table 2. The estimated band gap energy (Eg), conduction band edge (CB), and valence band edge (VB) for BiOCl and ZnIn2S4.
Table 2. The estimated band gap energy (Eg), conduction band edge (CB), and valence band edge (VB) for BiOCl and ZnIn2S4.
SemiconductorEg (eV)CB (eV)VB (eV)
BiOCl3.20−1.321.88
ZnIn2S42.61−1.361.25

Share and Cite

MDPI and ACS Style

Guo, L.; Han, X.; Zhang, K.; Zhang, Y.; Zhao, Q.; Wang, D.; Fu, F. In-Situ Construction of 2D/2D ZnIn2S4/BiOCl Heterostructure with Enhanced Photocatalytic Activity for N2 Fixation and Phenol Degradation. Catalysts 2019, 9, 729. https://doi.org/10.3390/catal9090729

AMA Style

Guo L, Han X, Zhang K, Zhang Y, Zhao Q, Wang D, Fu F. In-Situ Construction of 2D/2D ZnIn2S4/BiOCl Heterostructure with Enhanced Photocatalytic Activity for N2 Fixation and Phenol Degradation. Catalysts. 2019; 9(9):729. https://doi.org/10.3390/catal9090729

Chicago/Turabian Style

Guo, Li, Xuanxuan Han, Kailai Zhang, Yuanyuan Zhang, Qiang Zhao, Danjun Wang, and Feng Fu. 2019. "In-Situ Construction of 2D/2D ZnIn2S4/BiOCl Heterostructure with Enhanced Photocatalytic Activity for N2 Fixation and Phenol Degradation" Catalysts 9, no. 9: 729. https://doi.org/10.3390/catal9090729

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