Next Article in Journal
Variation of Nonlinear Refraction and Three-Photon Absorption of Indium–Tin Oxide Quantum Dot Thin Films and Solutions in Near Infrared Range
Previous Article in Journal
N-Octadecane Encapsulated by Assembled BN/GO Aerogels for Highly Improved Thermal Conductivity and Energy Storage Capacity
Previous Article in Special Issue
Controllable Doping Characteristics for WSxSey Monolayers Based on the Tunable S/Se Ratio
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Construction of 2D/2D ZnIn2S4-Based Bifunctional Photocatalysts for H2 Production and Simultaneous Degradation of Rhodamine B and Tetracycline

1
Jiangxi Key Laboratory of Surface Engineering, School of Materials and Energy, Jiangxi Science and Technology Normal University, Nanchang 330013, China
2
Key Laboratory of Novel Biomass-Based Environmental and Energy Materials in Petroleum and Chemical Industry, Key Laboratory of Green Chemical Engineering Process of Ministry of Education, Engineering Research Center of Phosphorus Resources Development and Utilization of Ministry of Education, Hubei Key Laboratory of Novel Reactor and Green Chemical Technology, School of Chemistry and Environmental Engineering, Wuhan Institute of Technology, Wuhan 430073, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(16), 2315; https://doi.org/10.3390/nano13162315
Submission received: 25 July 2023 / Revised: 10 August 2023 / Accepted: 10 August 2023 / Published: 12 August 2023
(This article belongs to the Special Issue 2D Semiconductor Nanomaterials and Heterostructures)

Abstract

:
A two-dimensional/two-dimensional (2D/2D) TiO2/ZnIn2S4 photocatalyst was reasonably proposed and constructed by a two-step oil bath-hydrothermal method. TiO2 nanosheets uniformly grown on the surface of ZnIn2S4 nanosheets and a synergetic effect between the TiO2 and ZnIn2S4 could highly contribute to improving the specific surface area and hydrophilicity of ZnIn2S4 as well as accelerating the separation and transfer of photon-generated e-h+ pairs, and thus enhancing the visible-light photocatalytic degradation and H2 evolution performance of ZnIn2S4. Rhodamine B (RhB) and tetracycline (TC) were simultaneously selected as the target pollutants for degradation in the work. The optimum photocatalytic RhB and TC degradation properties of TiO2/ZnIn2S4-10 wt% were almost 3.11- and 8.61-fold higher than that of pure ZnIn2S4, separately, while the highest photocatalytic hydrogen evolution rate was also observed in the presence of TiO2/ZnIn2S4-10wt% and 4.28-fold higher than that of ZnIn2S4. Moreover, the possible photocatalytic mechanisms for enhanced visible-light photocatalytic degradation and H2 evolution were investigated and proposed in detail. Our research results open an easy pathway for developing efficient bifunctional photocatalysts.

1. Introduction

Since the concept of sustainable development was proposed, the production of clean energy and the treatment of wastewater with persistent organic pollutants have attracted increasing attention from researchers [1,2,3,4]. Compared to conventional treatment methods, photocatalysis technology by semiconductors has some advantages of clean, easy operation and high efficiency, which is considered to be promising in the territory of alleviating energy shortages and environmental crises [5,6]. Numerous scholars have been endeavoring to probe newfashioned semiconductors photocatalysts with superior activity and good stability to achieve effective hydrogen production and pollutant degradation in the past few decades [7,8]. Among the semiconductors photocatalysts, ternary metal chalcogenide semiconductors, such as CuCo2S4, ZnIn2S4, and CaIn2S4, have obtained exceeding attention in the domain of photocatalysis research owing to the advantages of small band gaps, outstanding photoconversion capacity, and good stability [9,10,11].
ZnIn2S4, as an outstanding representative of ternary metal chalcogenides, possesses two-dimensional (2D) nanosheet morphologies, a narrow bandgap of ca. 2.4 eV, and good stability, and thus is recognized to be a suitable candidate for visible-light photocatalytic hydrogen production and pollutant degradation [12,13]. Nonetheless, pristine ZnIn2S4 tends to agglomerate and displays low separation efficiency of the photogenerated electron-hole pairs, which greatly restricts its photocatalytic property with hindering its application in the photocatalytic realm [14]. Therefore, it is urgently needed to surmount the drawbacks of pristine ZnIn2S4, and thus a series of modification strategies have been proposed. Among all of them, constructing heterojunctions with other semiconductor materials can immensely promote the separation of photogenerated electron-hole pairs, which has been demonstrated to be a productive modification strategy [15,16].
Among the semiconductor materials, TiO2, as a wide bandgap (Eg~3.2 eV) semiconductor material, is widely considered as an ideal candidate to fabricate heterojunction with ZnIn2S4 due to its excellent stability, nontoxicity, and low cost [17], which has been widely applied in photocatalytic H2 production [18], pollution degradation [19], CO2 reduction [20], and organic synthesis [21]. More importantly, the energy band position of ZnIn2S4 is above that of TiO2 allowing for photogenerated carriers transfer between ZnIn2S4 and TiO2, and thus coupling ZnIn2S4 and TiO2 contributes to addressing the shortfalls of ZnIn2S4 [22]. So far, TiO2/ZnIn2S4 heterojunctions with different morphologies, such as 2D/3D TiO2 nanosheets/ZnIn2S4 nanostructure [23,24], 3D ZnIn2S4 nanosheets/TiO2 nanobelts [25], and 1D TiO2 nanofibers/2D ZnIn2S4 nanosheet heterostructure [26], have been successfully fabricated with significantly improving the separation efficiency and lifetime of carriers, and thus boosting the photocatalytic activity of ZnIn2S4. It is widely believed that the 2D/2D structure with close contacts has potential advantages of large specific surface area, excellent light absorption ability, and effective charge separation efficiency [27,28,29]. It was revealed that due to the 2D/2D structure, Co3O4/ZnIn2S4 and TiO2/g-C3N4 photocatalysts showed efficient separation of photogenerated carriers, and thus obtaining enhanced photocatalytic properties [30,31]. Therefore, it is necessary to fabricate 2D/2D TiO2/ZnIn2S4 nanostructures and investigate the enhanced photocatalytic activity.
Enlightened by the aforementioned studies, we attempt to design and synthesize the TiO2/ZnIn2S4 nanocomposites with intimate contacted 2D/2D structure by growing TiO2 nanosheets on the surfaces of ZnIn2S4 nanosheets. The synergistic effect between TiO2 and ZnIn2S4 promoted the photogenerated carriers’ separation as well as enhanced specific surface area and hydrophilicity. As a result, the as-obtained composite photocatalysts showed significantly enhanced photocatalytic H2 production rate and pollution removal efficiency with excellent reusability. The charge separation and transfer mechanism on the contact interface of TiO2 and ZnIn2S4 for the superior photocatalytic performance was analyzed in-depth. This study provides a promising path for the construction of highly efficient photocatalysts for simultaneous application in energy- and environment-related areas.

2. Materials and Methods

2.1. Chemicals

HF (40% aqueous solution), tetraisopropyl titanate (TIPT, ≥95.0%), NaOH (≥99.0%), chromic chloride (CdCl3, ≥99.0%), and indium chloride (InCl3, ≥99.9%) were purchased from Macklin Biochemical Co., Ltd. (Shanghai, China). RhB (≥99.0%), TC (≥99.0%), Thioacetamide (TAA, ≥99.0%), anhydrous ethanol (≥99.7%), hydrochloric acid (HCl, 36.5%), and zinc chloride (ZnCl2, ≥98.0%) were commercially available from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Ethylenediamine tetraacetic acid disodium (EDTA-2Na, ≥99.5%), p-Benzoquinone (BQ, ≥99.5%), tertiary butyl alcohol (t-BuOH, ≥99.5%), and triethanolamine (TEOA, ≥98.0%) were supplied by Shanghai Zhanyun Chemical Co., Ltd. (Shanghai, China). All chemicals were utilized as received. The distilled water was obtained using Water Purification System.

2.2. Synthesis of ZnIn2S4 Nanosheets

ZnIn2S4 was prepared via an oil-bath process according to the former literature [12]. 272 mg of ZnCl2, 442 mg of InCl3, and 300 mg of TAA were dissolved into 50 mL of deionized water (pH = 2.5), and heated at 80 °C for 2 h. After cooling, ZnIn2S4 could be obtained by separating, washing, and drying under a vacuum at 60 °C overnight. Finally, 500 mg of ZnIn2S4 was dispersed into 100 mL of methanol with continuous ultrasound treatment.

2.3. Synthesis of TiO2/ZnIn2S4 Nanosheets

The specific reaction process was illustrated in Figure 1. Firstly, 10 mL of HF was slowly dropped into a 100 mL of Teflon-lined autoclave reactor containing 25 mL of tetrabutyl titanate and heated at 200 °C for 40 h. After cooling, the precipitates were thoroughly separated by centrifugation and then dried under vacuum at 60 °C for overnight. Subsequently, the precipitates (3.1 mg, 15.5 mg, 18.6 mg, and 31 mg) and ZnIn2S4 dispersion liquid (19.6 mL, 18 mL, 17.6 mL, and 16 mL) were added in 0.1 M NaOH solution under stirring for 24 h, separately, then washed with deionized water till the pH = 7 and dried at 100 °C in a vacuum drying chamber. Finally, TiO2/ZnIn2S4 composites with different weight percent of TiO2 (2 wt%, 10 wt%, 12 wt% and 20 wt%) were obtained via the procedure and marked as TiO2/ZnIn2S4-2 wt%, TiO2/ZnIn2S4-10 wt%, TiO2/ZnIn2S4-12 wt% and TiO2/ZnIn2S4-20 wt%, respectively. Further, the identical method was also utilized to synthesize blank TiO2 in the absence of ZnIn2S4.

2.4. Characterization

The crystal phases were investigated via an X-ray diffractometer (XRD, XRD-6100, Shimadzu, Kyoto, Japan) using Cu-Kα radiation (λ = 1.5406 Å). The morphologies and lattice properties were analyzed by scanning electron microscopy (SEM, Sigma, Carl Zeiss, Oberkochen, Germany) and transmission electron microscopy (TEM, JEM2100, JEOL, Kyoto, Japan). The specific surface areas were determined by the physical adsorption of N2 on a Micromeritics (ASAP 2020, Micromeritics, Atlanta, GA, USA) using the Brunauer–Emmett–Teller (BET) equation. The chemical state was analyzed by X-ray photoelectron spectroscopy (XPS, Thermo Scientific K-Alpha, Thermo Fisher, Waltham, MA, USA). The light absorption as well as charge separation and transfer efficiency were studied by ultraviolet-visible diffuse reflection spectroscopy (UV–vis DRS, Lambda 750, PerkinElmer, Waltham, USA), photoluminescence spectroscopy (PL, ZolixLSP-X500A, Zolix, Beijing, China), fluorescence lifetime spectrophotometer (C11367, Quantaurus-Tau, Hamamatsu, Japan), and three-electrode photoelectrochemical cell system (CHI660E, Chenghua, Shanghai, China). The water contact angles were measured by a contact angle meter (HARKE-SPCA, HARKE, Beijing, China). TOC analyzer (TOC-2000, Metash, Shanghai, China) was utilized to investigate the total organic carbon (TOC) of the residual solution.

2.5. Density Functional Theory (DFT) Calculation

To calculate the band gaps and work functions of TiO2 and ZnIn2S4, the model of TiO2 (1 × 2 × 1 supercell) and ZnIn2S4 (1 × 1 × 2 supercell) were first built and given in Figure S1. Subsequently, the calculations were performed by utilizing the Vienna ab initio Simulation Package (VASP), which implements the DFT with a generalized gradient approximation (GGA) and super-soft pseudopotential method. Calculations were carried out by utilizing the Predew–Burke–Ernzerhof (PBE) scheme. The electron wave functions were described by the projector augmented wave (PAW) method with a cutoff energy of 300 eV and a K-point of 2 × 3 × 2.

2.6. Photocatalytic Hydrogen Generation

The photocatalytic hydrogen experiments were conducted in an automatic online gas analysis system (Labsolar-6A, Perfectlight, Beijing, China). A Xenon lamp of 300 W (PLS-SXE 300C, λ > 420 nm) was employed to supply the visible-light source. During the process, 20 mg of as-fabricated photocatalysts were added into a 60 mL of mixed solution (50 mL deionized water and 10 mL of TEOA) without adding H2PtCl6, and then the reaction container was installed into the photocatalytic reaction instrument and the distance between the light source and the solution was about 16 cm. Ahead of starting the reaction, the entire installation was vacuumed to remove the air until the system pressure was beneath 1.0 Kpa. Then, turned on the light source and operated the program, automatic sampling every 60 min. In the whole reaction process, the temperature of circulating cooling water was always controlled at about 5 °C. Finally, the generated blended gas was transferred to gas chromatography (GC9790) equipped with a TCD detector (LabSolar-IIIAG, Perfectlight, Beijing, China) to further detect and calculate the production of hydrogen. To investigate the reusability of the binary heterostructure, recycled hydrogen production was carried out four times using TiO2/ZnIn2S4-10 wt% as the photocatalyst. The apparent quantum efficiencies (AQE) for H2 evolution of λ = 400, 420, and 500 nm were determined in a 75 mL Pyrex glass reactor. The apparent quantum efficiency (AQE) could be determined using the following equation:
AQE = 2 × the   number   of   H 2   evolved   molecules the   number   of   incident   photons × 100 %

2.7. Photodegradation Activity Evaluation

The visible-light photocatalytic degradation activity was evaluated by the degradation of fresh TC (10 mg/L) and RhB (30 mg/L) solution. The light source was provided by 500 W Xenon light (PLS-SXE 500C) equipped with a UV cutoff filter (λ > 420 nm). Firstly, 10 mg of photocatalyst was dispersed into the 50 mL of TC solution and 50 mL of RhB solution, respectively. Then, the above-mentioned solution was transferred to the photocatalytic reaction apparatus (XPA-7) and kept stirring in the dark for 60 min to obtain an adsorption-desorption equilibrium. After turning on the Xenon light, the photocatalytic degradation reaction was starting. At a given interval, a 4 mL aliquot of mixture was taken out utilizing a syringe with a needle and then filtrated using a 0.22 μm Millipore filter to obtain the residual solution, the concentration of which at the maximum absorption wavelength (355 nm for TC and 554 nm for RhB) was monitored using a PerkinElmer UV–vis spectrophotometer (Lambda 35). Moreover, the recycled photodegradation experiments were carried out four times using TiO2/ZnIn2S4-10 wt% as the photocatalyst at the same condition. Once the degradation experiment was over, the remaining sample in the beaker was immediately recycled by separation, washing, and drying for the next cyclic degradation experiment. The degradation efficiency (De %) was calculated by the equation:
De   % = C 0 C t C 0 × 100 %
In the formula, C0 and Ct denote the concentrations at the initial time and after each stage of degradation, separately. As for the trapping experiments, equal amounts (1.0 mM) of scavengers were added to the TC solution to capture active radicals.

3. Results and Discussions

The text continues here. The XRD patterns were used to recognize the phase composition and the structure of samples, and the results are exhibited in Figure 2. The diffraction peaks observed at 21.6°, 27.7°, 47.2°, 52.7°, and 55.6° can be well attributed to the (006), (102), (110), (116), and (022) crystal planes of hexagonal ZnIn2S4 (JCPDS No. 72-0773) [32]. For pure TiO2, a set of diffraction peaks located at 25.3° (101), 37.8° (004), 48.1° (200), 54.1° (105), 55.1° (211) and 62.8° (204) are consistent with the anatase TiO2 (JCPDS No. 21-1272) [33]. The diffraction peaks of ZnIn2S4 and TiO2 can be seen simultaneously in the XRD patterns of TiO2/ZnIn2S4 composites, the characteristic peak intensities of TiO2 generally increased, while the diffraction peaks of ZnIn2S4 gradually decrease with the increased content of TiO2. Further, no other peaks of the third (impurity) phase are detected in all XRD patterns, implying the successful construction of the TiO2/ZnIn2S4 composites.
SEM and TEM were utilized to observe the morphology of the catalysts. As displayed in Figure 3a, pristine TiO2 showed a 2D nanosheet structure with different sizes. As for pure ZnIn2S4, a nanoflower-like structure assembled by the large number of nanosheets can be seen in Figure 3b. After coupling ZnIn2S4 with TiO2, SEM (Figure 3c) and TEM (Figure 3d) images indicate TiO2 nanosheets grow on the surface of ZnIn2S4 nanosheets, forming an intimate 2D/2D contact interface between ZnIn2S4 and TiO2. The elemental distribution in the TiO2/ZnIn2S4 composite was analyzed using SEM-energy-dispersive X-ray spectroscopy (Figure S3), and the result confirmed the homogeneous coexistence of Ti, O, Zn, In, and S elements. Moreover, high-resolution TEM analysis was conducted to investigate the microstructure information of the TiO2/ZnIn2S4-10 wt% composite, and the result is depicted in Figure 3e, the lattice fringes with d spacings of 0.352 and 0.322 nm can be seen, which can be assigned to TiO2 (101) and ZnIn2S4 (102) facets, separately [34,35]. The above results further indicate the successful formation of the TiO2/ZnIn2S4 hybrid.
XPS was explored to be aware of the surface element composition and chemical state of the TiO2/ZnIn2S4-10 wt% composite. As shown in Figure 4a, the XPS survey spectrum of TiO2/ZnIn2S4 reveals the coexistence of Ti, O, S, Zn, and In elements, which is in keeping with the EDS test results. Figure 4b presents the XPS spectrum of O 1s, two characteristic peaks located at 530.6 and 531.92 eV can be attributed to the Ti-O bond and the –OH group, respectively [36]. The high-resolution XPS spectra of Ti 2p showed two characteristic peaks located at 458.34 and 463.89 eV (Figure 4c), assigning to Ti 2p3/2 and Ti 2p1/2, separately [37]. In the high-resolution S 2p spectrum (Figure 4d), the binding energies of 161.06 and 162.31 eV can be assigned to the S 2p3/2 and S 2p1/2, respectively, suggesting the occurrence of S2− [38]. In Figure 4e, the peaks centered at 444.30 and 452.40 eV are ascribed to In 3d5/2 and In 3d3/2, assigning to In3+ binding state [39]. As for Zn 2p (Figure 4f), the peaks centered at 1021.34 and 1044.36 eV ascribed to 2p3/2 and 2p1/2, respectively, which proves the existence of Zn2+ [40].
The specific surface area and water contact angles were measured to investigate the adsorption performance of photocatalysts, and the results are displayed in Figure 5. The nitrogen adsorption-desorption isotherms of TiO2, ZnIn2S4, and TiO2/ZnIn2S4-10 wt% showed type IV isotherms with the hysteresis loop of mesoporous structures (Figure 5a), the specific surface area of TiO2/ZnIn2S4-10 wt% was larger than that of TiO2 and ZnIn2S4. It can be seen that the average pore sizes are between 2 and 50 nm (Figure 5b), which further confirmed the formation of a mesoporous structure. Meanwhile, water contact angles of TiO2, TiO2/ZnIn2S4, and ZnIn2S4 were also measured to analyze the hydrophilicity and hydrophobicity of prepared materials. It can be observed from Figure 5c–e that contact angles of ZnIn2S4, TiO2/ZnIn2S4-10 wt% and TiO2 were, respectively, 72.9°, 15.6°, and 8.9°, manifesting the hydrophilicity of ZnIn2S4 can be improved by coupling with TiO2. These results illustrated that interface contact exists between pollutants and the photocatalysts owing to the enhanced specific surface area and hydrophilicity, and thus it is expected to obtain excellent photocatalytic performance.
The photoabsorptive behavior of TiO2, ZnIn2S4, and TiO2/ZnIn2S4-10 wt% was detected via UV–vis DRS spectra as shown in Figure 6a, the absorption wavelength of pristine ZnIn2S4 with steep edge was at approximate 560 nm in the visible-light areas, which presented favorable absorption capacity both in the visible and UV light, while the absorption edge of absolute TiO2 was located in about 406 nm. The photoabsorption ability of TiO2/ZnIn2S4-10 wt% exhibits a very close absorption profile with ZnIn2S4 with slightly diminished absorption and blue-shifted absorption edge, suggesting the introduction of TiO2 has a slight influence on the light absorption property of ZnIn2S4. As depicted in Figure 6b,c, the band gaps of TiO2 and ZnIn2S4 were calculated as 3.24 and 2.48 eV based on the equation: (αhν)1/n = A(hν−Eg) [41], which were roughly matched with the results of DFT calculation (Figure S3). Valence band XPS (VB-XPS) of TiO2 and ZnIn2S4 were also conducted to further understand the band structure of TiO2 and ZnIn2S4. As demonstrated in Figure 6d, the EVB-XPS values of pure TiO2 and ZnIn2S4 were 2.95 and 1.49 eV, respectively. Therefore, the VB potentials of the normal hydrogen electrode (EVB–NHE, pH = 7) of TiO2 and ZnIn2S4 were determined to be 2.71 and 1.25 eV based on the EVB vs. NHE = φ + EVB-XPS − 4.44, where φ is the work function (4.2 eV) of the XPS analyzer [42], while the ECB vs. NHE values of TiO2 and ZnIn2S4 could be computed as −0.53 and −1.23 eV using the equation: ECB = EVB − Eg [43]. Therefore, the overall band structure positions of TiO2 and ZnIn2S4 can be obtained and shown in Figure S4.
To uncover the positive influence of constructing heterojunction on the catalytic performances of ZnIn2S4, the separation and migration behaviors of photogenerated charges were deeply investigated. First, the PL spectra of ZnIn2S4 and TiO2/ZnIn2S4-10 wt% were measured to monitor the recombination process of photoinduced charge carriers. Generally, the dramatically reduced PL intensity is regarded as a signal of effective charge separation [44]. As shown in Figure 7a, TiO2/ZnIn2S4-10 wt% exhibited a lower PL intensity than ZnIn2S4, suggesting a higher separation efficiency of photogenerated carriers in TiO2/ZnIn2S4-10 wt% composite. The time-resolved photoluminescence (TRPL) spectra were also acquired to investigate the detailed information about the decay behavior of photogenerated carriers, and the results are shown in Figure 7b. The average fluorescence lifetime (τavg) of TiO2/ZnIn2S4-10 wt% (581 ps) was longer than pristine ZnIn2S4 (543 ps), implying that the coupling ZnIn2S4 with TiO2 can helpfully prevent the recombination of photoinduced carriers and obtain a longer fluorescence lifetime. To further clarify the enhanced photogenerated charge transfer and separation efficiency, the photoelectrochemical performance was characterized and analyzed by transient photocurrent responses and EIS tests. As demonstrated in Figure 7c, the TiO2/ZnIn2S4-10 wt% showed a higher photocurrent density than ZnIn2S4, and the average photocurrent density of TiO2/ZnIn2S4-10 wt% was raised to be 2.11 mA·cm−2, approximately 1.5-fold larger than that of pristine ZnIn2S4 (1.39 mA·cm−2), implying that the construction of TiO2/ZnIn2S4 composite can promote the photoexcited charge carrier transfer. Furthermore, the EIS plot of TiO2/ZnIn2S4-10 wt% composite exhibited a smaller semicircle than pristine ZnIn2S4, as observed in Figure 7d, manifesting a lesser electric resistance and more efficient charge transfer process existing in the TiO2/ZnIn2S4-10 wt% composite. These optical and photoelectrochemical properties demonstrated that the formation of TiO2/ZnIn2S4 heterojunction was capable of elevating the separation and transfer efficiency of photogenerated carriers, thus obtaining the enhanced photocatalytic performance.
The visible-light photocatalytic H2 generation activities of TiO2, ZnIn2S4, and TiO2/ZnIn2S4 composites were evaluated in the presence of TEOA sacrificial reagent. As shown in Figure 8a, pure ZnIn2S4 possessed a low photocatalytic performance due to the high recombination rate of photoexcited charge carriers and photocorrosion, while pristine TiO2 had almost no catalytic activity, which could be attributed to its wide bandgap [45]. Notably, the photoactivity of ZnIn2S4 was gradually improved along with the introduction of TiO2. Among all composites, the TiO2/ZnIn2S4-10 wt% composite showed the optimal H2 rate of 650 μmol/h/g, which was 4.28-fold higher than that of pristine ZnIn2S4 (Figure 8b). The recycling tests were also conducted in the same reaction condition to investigate the durability performance of TiO2/ZnIn2S4-10 wt% photocatalyst. As is demonstrated in Figure 8c, the H2 production amount throughout four successive cycles barely changed. Moreover, the crystal structure and morphology of TiO2/ZnIn2S4-10 wt% showed no noticeable changes by comparing the XRD pattern (Figure S5a) or SEM image (Figure S5b) of a used sample with the fresh sample. These test results manifested that the TiO2/ZnIn2S4-10 wt% composite possesses good photocatalytic stability. To further clarify the driving force in the photocatalytic process, the AQEs of TiO2/ZnIn2S4-10 wt% photocatalyst at 400, 420, and 500 nm were calculated as 1.3, 1.1, and 0.1%, respectively (Figure 8d), which exhibits a similar trend with the adsorption spectrum, indicating that the H2 production reaction is a photocatalytic driven process.
To confirm the performance multiformity of the as-prepared samples, the photocatalytic degradation capacities of all samples were also investigated by using colorless TC and colored RhB as the simulated organic pollutants. The TC and RhB photodegradation curves of TiO2, ZnIn2S4, and TiO2/ZnIn2S4 composites were illustrated in Figure 9a and Figure S6a, respectively. Almost no changes in the concentration of TC and RhB were noticed in the absence of a catalyst, suggesting that the self-degradation process could be ignored. The pure TiO2 displayed weak degradation activities within 60 min, while ZnIn2S4 showed high degradation activities than TiO2 due to a wider visible-light response range. With respect to the TiO2/ZnIn2S4 composites, all composites showed better photocatalytic performance than TiO2 and ZnIn2S4. Among them, TiO2/ZnIn2S4-10 wt% possessed the optimum performance, and almost 95% of TC and 93% of RhB could be degraded. The photocatalytic activity was enhanced when the mass content of TiO2 increased from 2% to 10%, then decreased as TiO2 content further increased to 12% or even more, which may be attributed to excessive TiO2 shielding the light absorption. To obtain the reaction rate constant “k”, the photodegradation curves were further kinetically fitted by using the pseudo-first-order equation: −ln (C/C0) = kt, the results were displayed in Figure 9b and Figure S6b, separately. The k value of TiO2/ZnIn2S4-10 wt% composite was highest compared with other samples and was up to 0.04115 min−1 for TC and 0.04168 min−1 for RhB, which was almost 111 and 190 fold that of pure TiO2, and 26 and 6.65 fold that of individual ZnIn2S4. Meanwhile, the mineralization capacities of all kinds of photocatalysts were investigated by TOC measurement. As demonstrated in Figure 9c and Figure S6c, the TOC removal efficiencies of TiO2/ZnIn2S4 composites distinctly overtopped TiO2 and ZnIn2S4 under the irradiation of visible light. Among them, the TiO2/ZnIn2S4-10 wt% composite showed the highest TOC removal efficiency (83.5% for TC and 85.6 for RhB), which formed the correspondence with its doughty photocatalytic degradation abilities, and confirmed that the TiO2/ZnIn2S4 had high mineralization capacities. To determine the reusability of TiO2/ZnIn2S4-10 wt% in the photocatalytic process, the photocatalytic cycle experiments were performed to investigate the reusable performance. As shown in Figure 9d and Figure S6d, the photodegradation efficiency scarcely had changed after undergoing four consecutive cycles. In addition, the XRD pattern (Figure S7a,c) and SEM image (Figure S7b,d) of TiO2/ZnIn2S4-10 wt% illuminated the crystal structure and morphology of TiO2/ZnIn2S4-10 wt% before and after photodegradation cycling remained unchanged. The results demonstrated the splendid degradation stability of TiO2/ZnIn2S4-10 wt% during the photocatalytic process.
The work functions (Φ) were calculated to investigate the route of charge transfer at the contact interface of ZnIn2S4 and TiO2, and the results are given in Figure 10a. It was observed that the Φ of ZnIn2S4 is lower than that of TiO2, and thus the photoinduced electrons could transfer from ZnIn2S4 to TiO2 when ZnIn2S4 and TiO2 came in contact to construct a heterojunction. Subsequently, EDTA-2Na, t-BuOH, and BQ were selected in sequence as the scavengers of h+, OH, and O2 to further identify the roles of active species during the photodegradation process. As recorded in Figure 10b, varying degrees of photocatalytic activity suppression were observed after sacrificial agents were added with an order BQ > EDTA-2Na > t-BuOH, indicating O2 and h+ are main and secondary active substances, separately, while OH has minimal impact on the photocatalytic reactions.
The possible mechanisms for boosting photocatalytic pollutant degradation and H2 production performances of TiO2/ZnIn2S4 composites were proposed and illustrated in Figure 11 based on the aforementioned discussion. When TiO2 nanosheets grew on the surface of ZnIn2S4 nanosheets, a closed contact interface was formed between ZnIn2S4 and TiO2. Under visible-light irradiation, TiO2 could not absorb visible light due to its large band gap energy, while the electrons on the VB of ZnIn2S4 could be easily excited to its CB and generate electron-hole pairs because of the small band gap energy. According to the DFT calculated results of work functions, the electrons would transfer from the CB of ZnIn2S4 to that of TiO2, while h+ left on the VB of ZnIn2S4, which leads to the spatial separation of electrons and holes with higher redox powers. For H2 production, the photogenerated e could easily reduce the surface-adsorbed protons to H2 with h+ being consumed by TEOA. Different from the H2 production process, electrons on the CB of TiO2 would firstly react with O2 to obtain O2 in the photodegradation process. Subsequently, the h+ and O2 participated in the photodegradation reaction due to their strong oxidation ability. As a result, the improved separation efficiency of photoinduced charge carriers would provide more carriers to participate in the photocatalytic reaction process to acquire enhanced photocatalytic performance.

4. Conclusions

In summary, a 2D/2D heterojunction consisting of ZnIn2S4 nanosheets and TiO2 nanosheets was fabricated by a facile two-step synthesis method for photocatalytic H2 evolution and pollutant degradation. The small TiO2 nanosheets deposited on the surface of large ZnIn2S4 nanosheets, resulting in the formation of the close 2D/2D heterointerface contact, which contributes to providing sufficient and short paths for the separation and transfer of photoinduced charge. All TiO2/ZnIn2S4 heterojunction photocatalysts possess higher photocatalytic activities than pure ZnIn2S4 and TiO2. Among them, the TiO2/ZnIn2S4-10 wt% photocatalyst exhibits optimal H2 evolution rate (650 μmol/h/g) and pollution degradation efficiencies (95% for TC and 93% for RhB) with excellent photocatalytic stability during four consecutive test cycles. The enhanced photocatalytic properties are believed to have originated from the accelerated charges separation and transfer as well as enhanced specific surface area and hydrophilicity. This work provides a practical strategy for preparing ZnIn2S4-based heterojunctions to act as highly efficient bifunctional photocatalysts for energy and environmental application.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13162315/s1, Figure S1: structure models; Figure S2: The elemental mapping images; Figure S3: Calculated band structures; Figure S4: The energy band structure; Figure S5: XRD patterns and SEM image in the recycled photocatalytic H2 development; Figure S6: Photocatalytic degradation rate, pseudo-first-order kinetics fitted curves, TOC removal rate and maintenance of catalytic performance; Figure S7: XRD patterns and SEM images in the recycling photocatalytic degradation.

Author Contributions

Y.C.: investigation, methodology, data curation, and writing—original draft preparation. L.Z.: data curation, validation, and formal analysis. Y.S. and J.L.: software, investigation. J.X.: resources, supervision, and writing—review and editing. L.Q. and X.X.: investigation and methodology. D.M.: validation and funding acquisition. P.L.: conceptualization, methodology, funding acquisition, and writing—review and editing. S.D.: project administration, supervision, and funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the National Natural Science Foundation of China (grant number 41763020) and the Natural Science Foundation of Jiangxi Province (grant numbers 20212BAB204020, 20212BAB204018, and 20202BABL214040).

Data Availability Statement

The authors confirm that the data supporting the findings of this study are available within the article. Derived data supporting the findings of this study are available on request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, D.; Li, Y.; Wen, L.S.; Xi, J.B.; Liu, P.; Hansen, T.W.; Li, P. Ni-Pd-incorporated Fe3O4 yolk-shelled nanospheres as efficient magnetically recyclable catalysts for reduction of N-containing unsaturated compounds. Catalysts 2023, 13, 190. [Google Scholar] [CrossRef]
  2. Tan, H.; Li, J.L.; He, M.; Li, J.Y.; Zhi, D.; Qin, F.Z.; Zhang, C. Global evolution of research on green energy and environmental technologies: A bibliometric study. J. Environ. Manag. 2021, 297, 113382. [Google Scholar] [CrossRef]
  3. Xu, J.W.; Zheng, X.L.; Feng, Z.P.; Lu, Z.Y.; Zhang, Z.W.; Huang, W.; Li, Y.B.; Vuckovic, D.; Li, Y.Q.; Dai, S.; et al. Organic wastewater treatment by a single-atom catalyst and electrolytically produced H2O2. Nat. Sustain. 2021, 4, 233–241. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, G.P.; Wu, H.; Chen, D.Y.; Li, N.J.; Xu, Q.F.; Li, H.; He, J.H.; Lu, J.M. A mini-review on ZnIn2S4-based photocatalysts for energy and environmental application. Green Energy Environ. 2022, 7, 176–204. [Google Scholar] [CrossRef]
  5. Wang, J.; Sun, S.J.; Zhou, R.; Li, Y.Z.; He, Z.T.; Ding, H.; Chen, D.M.; Ao, W.H. A review: Synthesis, modification and photocatalytic applications of ZnIn2S4. J. Mater. Sci. Technol. 2021, 78, 1–19. [Google Scholar] [CrossRef]
  6. Liu, Q.L.; Zhao, Z.Y.; Zhao, R.D.; Yi, J.H. Fundamental properties of delafossite CuFeO2 as photocatalyst for solar energy conversion. J. Alloys Compd. 2020, 819, 153032. [Google Scholar] [CrossRef]
  7. Xiao, Y.; Wang, H.; Jiang, Y.H.; Zhang, W.L.; Zhang, J.M.; Wu, X.Y.; Liu, Z.C.; Deng, W. Hierarchical Sb2S3/ZnIn2S4 core-shell heterostructure for highly efficient photocatalytic hydrogen production and pollutant degradation. J. Colloid Interf. Sci. 2022, 623, 109–123. [Google Scholar] [CrossRef]
  8. Wei, Z.D.; Liu, J.Y.; Shangguan, W.F. A review on photocatalysis in antibiotic wastewater: Pollutant degradation and hydrogen production. Chinese J. Catal. 2020, 41, 1440–1450. [Google Scholar] [CrossRef]
  9. Guo, Y.C.; Yan, B.G.; Deng, F.; Shao, P.H.; Zou, J.P.; Luo, X.B.; Zhang, S.Q.; Li, X.B. Lattice expansion boosting photocatalytic degradation performance of CuCo2S4 with an inherent dipole moment. Chinese Chem. Lett. 2023, 34, 107468. [Google Scholar] [CrossRef]
  10. Wang, L.B.; Cheng, B.; Zhang, L.Y.; Yu, J.G. In situ irradiated XPS investigation on S-Scheme TiO2@ZnIn2S4 photocatalyst for efficient photocatalytic CO2 reduction. Small 2021, 17, 2103447. [Google Scholar] [CrossRef]
  11. Zhuge, Z.H.; Liu, X.J.; Chen, T.Q.; Gong, Y.Y.; Li, C.; Niu, L.Y.; Xu, S.Q.; Xu, X.T.; Alothman, Z.A.; Sun, C.Q.; et al. Highly efficient photocatalytic degradation of different hazardous contaminants by CaIn2S4-Ti3C2Tx Schottky heterojunction: An experimental and mechanism study. Chem. Eng. J. 2021, 421, 127838. [Google Scholar] [CrossRef]
  12. Chen, Y.; Zhu, L.Z.; Li, J.Z.; Qiu, L.F.; Zhou, C.H.; Xu, X.; Shen, Y.L.; Xi, J.B.; Men, D.D.; Li, P.; et al. Coupling ZnIn2S4 nanosheets with MoS2 hollow nanospheres as visible-light-active bifunctional photocatalysts for enhancing H2 evolution and RhB degradation. Inorg. Chem. 2023, 62, 7111–7122. [Google Scholar] [CrossRef] [PubMed]
  13. Zhao, H.; Jian, L.; Gong, M.; Jing, M.Y.; Li, H.X.; Mao, Q.Y.; Lu, T.B.; Guo, Y.X.; Ji, R.; Chi, W.W.; et al. Transition-metal-based cocatalysts for photocatalytic water splitting. Small Struct. 2022, 3, 2100229. [Google Scholar] [CrossRef]
  14. Yang, Z.F.; Xia, X.N.; Yang, W.W.; Wang, L.L.; Liu, Y.T. Photothermal effect and continuous hot electrons injection synergistically induced enhanced molecular oxygen activation for efficient selective oxidation of benzyl alcohol over plasmonic W18O49/ZnIn2S4 photocatalyst. Appl. Catal. B: Environ. 2021, 299, 120675. [Google Scholar] [CrossRef]
  15. Xia, M.Y.; Yan, X.Q.; Li, H.; Wells, N.; Yang, G.D. Well-designed efficient charge separation in 2D/2D N doped La2Ti2O7/ZnIn2S4 heterojunction through band structure/morphology regulation synergistic effect. Nano Energy 2020, 78, 105401. [Google Scholar] [CrossRef]
  16. Hao, C.C.; Tang, Y.B.; Shi, W.L.; Chen, F.Y.; Guo, F. Facile solvothermal synthesis of a Z-Scheme 0D/3D CeO2/ZnIn2S4 heterojunction with enhanced photocatalytic performance under visible light irradiation. Chem. Eng. J. 2021, 409, 128168. [Google Scholar] [CrossRef]
  17. Liu, H.; Li, J.Z.; Li, P.; Zhang, G.Z.; Xu, X.; Zhang, H.; Qiu, L.F.; Qi, H.; Duo, S.W. In-situ construction of 2D/3D ZnIn2S4/TiO2 with enhanced photocatalytic performance. Acta Chim. Sinica 2021, 79, 1293–1301. [Google Scholar] [CrossRef]
  18. Geng, Y.L.; Zou, X.L.; Lu, Y.N.; Wang, L. Fabrication of the SnS2/ZnIn2S4 heterojunction for highly efficient visible light photocatalytic H2 evolution. Int. J. Hydrogen Energy 2022, 47, 11520–11527. [Google Scholar] [CrossRef]
  19. Yang, R.J.; Chen, Q.Q.; Ma, Y.Y.; Zhu, R.S.; Fan, Y.Y.; Huang, J.Y.; Niu, H.N.; Dong, Y.; Li, D.; Zhang, Y.F.; et al. Highly efficient photocatalytic hydrogen evolution and simultaneous formaldehyde degradation over Z-scheme ZnIn2S4-NiO/BiVO4 hierarchical heterojunction under visible light irradiation. Chem. Eng. J. 2021, 423, 130164. [Google Scholar] [CrossRef]
  20. Du, J.; Shi, H.N.; Wu, J.M.; Li, K.Y.; Song, C.S.; Guo, X.W. Interface and defect engineering of a hollow TiO2@ZnIn2S4 heterojunction for highly enhanced CO2 photoreduction activity. ACS Sustainable Chem. Eng. 2023, 11, 2531–2540. [Google Scholar] [CrossRef]
  21. Wen, L.S.; Wang, D.; Xi, J.B.; Tian, F.; Liu, P.; Bai, Z.W. Heterometal modified Fe3O4 hollow nanospheres as efficient catalysts for organic transformations. J. Catal. 2022, 413, 779–785. [Google Scholar] [CrossRef]
  22. Huang, K.L.; Li, C.H.; Meng, X.C. In-situ construction of ternary Ti3C2 MXene@TiO2/ZnIn2S4 composites for highly efficient photocatalytic hydrogen evolution. J. Colloid Interf. Sci. 2020, 580, 669–680. [Google Scholar] [CrossRef] [PubMed]
  23. Liu, T.Y.; Wang, C.X.; Wang, W.; Xu, P.; Sun, X.N.; Zhang, J.T. The enhanced performance of Cr(VI) photoreduction and antibiotic removal on 2D/3D TiO2/ZnIn2S4 nanostructures. Ceram. Int. 2021, 47, 17015–17022. [Google Scholar] [CrossRef]
  24. Tan, Q.Y.; Li, K.N.; Li, Q.; Ding, Y.B.; Fan, J.J.; Xu, Z.H.; Lv, K.L. Photosensitization of TiO2 nanosheets with ZnIn2S4 for enhanced visible photocatalytic activity toward hydrogen production. Mater. Today Chem. 2022, 26, 101114. [Google Scholar] [CrossRef]
  25. 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: Environ. 2017, 219, 611–618. [Google Scholar] [CrossRef]
  26. Li, J.M.; Wu, C.C.; Li, J.; Dong, B.H.; Zhao, L.; Wang, S.M. 1D/2D TiO2/ZnIn2S4 S-scheme heterojunction photocatalyst for efficient hydrogen evolution. Chinese J. Catal. 2022, 43, 339–349. [Google Scholar] [CrossRef]
  27. Fu, J.W.; Xu, Q.L.; Low, J.X.; Jiang, C.J.; Yu, J.G. Ultrathin 2D/2D WO3/g-C3N4 step-scheme H2-production photocatalyst. Appl. Catal. B: Environ. 2019, 243, 556–565. [Google Scholar] [CrossRef]
  28. Hou, H.L.; Zeng, X.K.; Zhang, X.W. 2D/2D heterostructured photocatalyst: Rational design for energy and environmental applications. Sci. China Mater. 2020, 63, 2119–2152. [Google Scholar] [CrossRef] [Green Version]
  29. Mamiyev, Z.; Balayeva, N.O. Metal sulfide photocatalysts for hydrogen generation: A review of recent advances. Catalysts 2022, 12, 1316. [Google Scholar] [CrossRef]
  30. Zhang, G.P.; Li, X.X.; Wang, M.M.; Li, X.Q.; Wang, Y.R.; Huang, S.T.; Chen, D.Y.; Li, N.J.; Xu, Q.F.; Li, H.; et al. 2D/2D hierarchical Co3O4/ZnIn2S4 heterojunction with robust built-in electric field for efficient photocatalytic hydrogen evolution. Nano Res. 2023, 16, 6134–6141. [Google Scholar] [CrossRef]
  31. Zhang, Y.W.; Xu, J.S.; Mei, J.; Sarina, S.; Wu, Z.Y.; Liao, T.; Yan, C.; Sun, Z.Q. Strongly interfacial-coupled 2D-2D TiO2/g-C3N4 heterostructure for enhanced visible-light induced synthesis and conversion. J. Hazard. Mater. 2020, 394, 122529. [Google Scholar] [CrossRef]
  32. Qian, H.X.; Liu, Z.F.; Guo, Z.G.; Ruan, M.N.; Ma, J.L. Hexagonal phase/cubic phase homogeneous ZnIn2S4 n-n junction photoanode for efficient photoelectrochemical water splitting. J. Alloys Compd. 2020, 830, 154639. [Google Scholar] [CrossRef]
  33. Chen, Y.; Liu, H.; Hu, L.; Shen, Y.L.; Qiu, L.F.; Zhu, L.Z.; Shi, Q.X.; Xu, X.; Li, P.; Duo, S.W. Highly efficient visible-light photocatalytic performance of MOFs-derived TiO2 via heterojunction construction and oxygen vacancy engineering. Chem. Phys. Lett. 2023, 815, 140365. [Google Scholar] [CrossRef]
  34. Li, H.; Li, Y.H.; Wang, X.T.; Hou, B.R. 3D ZnIn2S4 nanosheets/TiO2 nanotubes as photoanodes for photocathodic protection of Q235 CS with high efficiency under visible light. J. Alloys Compd. 2019, 771, 892–899. [Google Scholar] [CrossRef]
  35. Sun, B.J.; Bu, J.Q.; Du, Y.C.; Chen, X.Y.; Li, Z.Z.; Zhou, W. O, S-dual-vacancy defects mediated efficient charge separation in ZnIn2S4/black TiO2 heterojunction hollow spheres for boosting photocatalytic hydrogen production. ACS Appl. Mater. Interfaces 2021, 13, 37545–37552. [Google Scholar] [CrossRef] [PubMed]
  36. Liu, S.Z.; Zhang, Y.C. Synthesis of CPVC-modified SnS2/TiO2 composite with improved visible light-driven photocatalysis. Mater. Res. Bull. 2021, 135, 111125. [Google Scholar] [CrossRef]
  37. Deng, Z.Q.; Li, L.; Ren, Y.C.; Ma, C.Q.; Liang, J.; Dong, K.; Liu, Q.; Luo, Y.L.; Li, T.S.; Tang, B.; et al. Highly efficient two-electron electroreduction of oxygen into hydrogen peroxide over Cu-doped TiO2. Nano Res. 2022, 15, 3880–3885. [Google Scholar] [CrossRef]
  38. Qiu, J.H.; Li, M.; Xu, J.; Zhang, X.F.; Yao, J.F. Bismuth sulfide bridged hierarchical Bi2S3/BiOCl@ZnIn2S4 for efficient photocatalytic Cr(VI) reduction. J. Hazard. Mater. 2020, 389, 121858. [Google Scholar] [CrossRef] [PubMed]
  39. 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]
  40. Xu, L.Z.; Deng, X.Y.; Li, Z.H. Photocatalytic splitting of thiols to produce disulfides and hydrogen over PtS/ZnIn2S4 nanocomposites under visible light. Appl. Catal. B: Environ. 2018, 234, 50–55. [Google Scholar] [CrossRef]
  41. Liu, C.; Ma, J.; Zhang, F.J.; Wang, Y.R.; Kong, C. Facile formation of Mo-vacancy defective MoS2/CdS nanoparticles enhanced efficient hydrogen production. Colloids Surf. A Physicochem. Eng. Asp. 2022, 643, 128743. [Google Scholar] [CrossRef]
  42. Cheng, S.; Qi, M.L.; Li, W.; Sun, W.Y.; Li, M.Q.; Lin, J.Y.; Bai, X.; Sun, Y.; Dong, B.; Wang, L. Dual-responsive nanocomposites for synergistic antibacterial therapies facilitating bacteria-infected wound healing. Adv. Healthc. Mater. 2023, 12, 2202652. [Google Scholar] [CrossRef] [PubMed]
  43. Shen, H.D.; Yang, C.M.; Xue, W.W.; Hao, L.D.; Wang, D.J.; Fu, F.; Sun, Z.Y. Construction of ternary bismuth-based heterojunction by using (BiO)2CO3 as electron bridge for highly efficient degradation of phenol. Chem. Eur. J. 2023, 29, e202300748. [Google Scholar] [CrossRef] [PubMed]
  44. Zhao, G.; Hao, S.H.; Guo, J.H.; Xing, Y.P.; Zhang, L.; Xu, X.J. Design of p-n homojunctions in metal-free carbon nitride photocatalyst for overall water splitting. Chinese J. Catal. 2021, 42, 501–509. [Google Scholar] [CrossRef]
  45. Liu, J.F.; Wang, P.; Fan, J.J.; Yu, H.G.; Yu, J.G. In situ synthesis of Mo2C nanoparticles on graphene nanosheets for enhanced photocatalytic H2-production activity of TiO2. ACS Sustain. Chem. Eng. 2021, 9, 3828–3837. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration for the formation process of TiO2/ZnIn2S4 composites.
Figure 1. Schematic illustration for the formation process of TiO2/ZnIn2S4 composites.
Nanomaterials 13 02315 g001
Figure 2. XRD patterns of the as-prepared ZnIn2S4, TiO2, and TiO2/ZnIn2S4 composites.
Figure 2. XRD patterns of the as-prepared ZnIn2S4, TiO2, and TiO2/ZnIn2S4 composites.
Nanomaterials 13 02315 g002
Figure 3. SEM images of TiO2 (a), ZnIn2S4 (b), and TiO2/ZnIn2S4-10wt% (c); TEM image (d) and HR-TEM image (e) of TiO2/ZnIn2S4-10 wt%.
Figure 3. SEM images of TiO2 (a), ZnIn2S4 (b), and TiO2/ZnIn2S4-10wt% (c); TEM image (d) and HR-TEM image (e) of TiO2/ZnIn2S4-10 wt%.
Nanomaterials 13 02315 g003
Figure 4. XPS spectra of as-synthesized TiO2/ZnIn2S4-10 wt% heterostructure: (a) survey scan, (b) O 1s, (c) Ti 2p, (d) S 2p, (e) In 3d, and (f) Zn 2p.
Figure 4. XPS spectra of as-synthesized TiO2/ZnIn2S4-10 wt% heterostructure: (a) survey scan, (b) O 1s, (c) Ti 2p, (d) S 2p, (e) In 3d, and (f) Zn 2p.
Nanomaterials 13 02315 g004
Figure 5. Nitrogen adsorption-desorption isotherms (a) and the corresponding pore size distribution plots (b) of bare TiO2, ZnIn2S4, and TiO2/ZnIn2S4-10 wt% samples; Water contact angles of ZnIn2S4 (c), TiO2/ZnIn2S4-10 wt% (d), and TiO2 (e).
Figure 5. Nitrogen adsorption-desorption isotherms (a) and the corresponding pore size distribution plots (b) of bare TiO2, ZnIn2S4, and TiO2/ZnIn2S4-10 wt% samples; Water contact angles of ZnIn2S4 (c), TiO2/ZnIn2S4-10 wt% (d), and TiO2 (e).
Nanomaterials 13 02315 g005
Figure 6. (a) UV–vis DRS of TiO2, ZnIn2S4 and TiO2/ZnIn2S4-10 wt%; energy bandgap of (b) TiO2 and (c) ZnIn2S4; (d) VB-XPS of TiO2 and ZnIn2S4.
Figure 6. (a) UV–vis DRS of TiO2, ZnIn2S4 and TiO2/ZnIn2S4-10 wt%; energy bandgap of (b) TiO2 and (c) ZnIn2S4; (d) VB-XPS of TiO2 and ZnIn2S4.
Nanomaterials 13 02315 g006
Figure 7. PL spectra (a), TRPL curves (b), transient photocurrent responses (c), and EIS plots (d) over bare ZnIn2S4 and TiO2/ZnIn2S4-10 wt% samples.
Figure 7. PL spectra (a), TRPL curves (b), transient photocurrent responses (c), and EIS plots (d) over bare ZnIn2S4 and TiO2/ZnIn2S4-10 wt% samples.
Nanomaterials 13 02315 g007
Figure 8. (a) Photocatalytic HER activity (a) and H2 evolution rates (b) of all samples. (c) Cycling experiments of TiO2/ZnIn2S4-10 wt%. (d) AQEs and DRS spectrum of TiO2/ZnIn2S4-10 wt%.
Figure 8. (a) Photocatalytic HER activity (a) and H2 evolution rates (b) of all samples. (c) Cycling experiments of TiO2/ZnIn2S4-10 wt%. (d) AQEs and DRS spectrum of TiO2/ZnIn2S4-10 wt%.
Nanomaterials 13 02315 g008
Figure 9. Photocatalytic degradation rate (a), the pseudo-first-order kinetics fitted curves (b), and TOC removal rate (c) of TC over all samples; maintenance of catalytic performance of TiO2/ZnIn2S4-10 wt% (d).
Figure 9. Photocatalytic degradation rate (a), the pseudo-first-order kinetics fitted curves (b), and TOC removal rate (c) of TC over all samples; maintenance of catalytic performance of TiO2/ZnIn2S4-10 wt% (d).
Nanomaterials 13 02315 g009
Figure 10. (a) Work functions of the TiO2 (001) and ZnIn2S4 (001) surfaces; (b) photocatalytic degradation performance of TC on TiO2/ZnIn2S4-10 wt% with various scavengers.
Figure 10. (a) Work functions of the TiO2 (001) and ZnIn2S4 (001) surfaces; (b) photocatalytic degradation performance of TC on TiO2/ZnIn2S4-10 wt% with various scavengers.
Nanomaterials 13 02315 g010
Figure 11. Proposed mechanisms for the photocatalytic reaction on the TiO2/ZnIn2S4 composite.
Figure 11. Proposed mechanisms for the photocatalytic reaction on the TiO2/ZnIn2S4 composite.
Nanomaterials 13 02315 g011
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

Chen, Y.; Zhu, L.; Shen, Y.; Liu, J.; Xi, J.; Qiu, L.; Xu, X.; Men, D.; Li, P.; Duo, S. Facile Construction of 2D/2D ZnIn2S4-Based Bifunctional Photocatalysts for H2 Production and Simultaneous Degradation of Rhodamine B and Tetracycline. Nanomaterials 2023, 13, 2315. https://doi.org/10.3390/nano13162315

AMA Style

Chen Y, Zhu L, Shen Y, Liu J, Xi J, Qiu L, Xu X, Men D, Li P, Duo S. Facile Construction of 2D/2D ZnIn2S4-Based Bifunctional Photocatalysts for H2 Production and Simultaneous Degradation of Rhodamine B and Tetracycline. Nanomaterials. 2023; 13(16):2315. https://doi.org/10.3390/nano13162315

Chicago/Turabian Style

Chen, Yue, Liezhen Zhu, Youliang Shen, Jing Liu, Jiangbo Xi, Lingfang Qiu, Xun Xu, Dandan Men, Ping Li, and Shuwang Duo. 2023. "Facile Construction of 2D/2D ZnIn2S4-Based Bifunctional Photocatalysts for H2 Production and Simultaneous Degradation of Rhodamine B and Tetracycline" Nanomaterials 13, no. 16: 2315. https://doi.org/10.3390/nano13162315

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